Anda di halaman 1dari 220

NCHRP

REPORT 782

Proposed Guideline
for Reliability-Based
Bridge Inspection Practices

NATIONAL
COOPERATIVE
HIGHWAY
RESEARCH
PROGRAM

TRANSPORTATION RESEARCH BOARD 2014 EXECUTIVE COMMITTEE*


OFFICERS
Chair: Kirk T. Steudle, Director, Michigan DOT, Lansing
Vice Chair: Daniel Sperling, Professor of Civil Engineering and Environmental Science and Policy; Director, Institute of Transportation Studies,
University of California, Davis
Executive Director: Robert E. Skinner, Jr., Transportation Research Board

MEMBERS
Victoria A. Arroyo, Executive Director, Georgetown Climate Center, and Visiting Professor, Georgetown University Law Center, Washington, DC
Scott E. Bennett, Director, Arkansas State Highway and Transportation Department, Little Rock
Deborah H. Butler, Executive Vice President, Planning, and CIO, Norfolk Southern Corporation, Norfolk, VA
James M. Crites, Executive Vice President of Operations, Dallas/Fort Worth International Airport, TX
Malcolm Dougherty, Director, California Department of Transportation, Sacramento
A. Stewart Fotheringham, Professor and Director, Centre for Geoinformatics, School of Geography and Geosciences, University of St. Andrews,
Fife, United Kingdom
John S. Halikowski, Director, Arizona DOT, Phoenix
Michael W. Hancock, Secretary, Kentucky Transportation Cabinet, Frankfort
Susan Hanson, Distinguished University Professor Emerita, School of Geography, Clark University, Worcester, MA
Steve Heminger, Executive Director, Metropolitan Transportation Commission, Oakland, CA
Chris T. Hendrickson, Duquesne Light Professor of Engineering, Carnegie Mellon University, Pittsburgh, PA
Jeffrey D. Holt, Managing Director, Bank of Montreal Capital Markets, and Chairman, Utah Transportation Commission, Huntsville, Utah
Gary P. LaGrange, President and CEO, Port of New Orleans, LA
Michael P. Lewis, Director, Rhode Island DOT, Providence
Joan McDonald, Commissioner, New York State DOT, Albany
Abbas Mohaddes, President and CEO, Iteris, Inc., Santa Ana, CA
Donald A. Osterberg, Senior Vice President, Safety and Security, Schneider National, Inc., Green Bay, WI
Steven W. Palmer, Vice President of Transportation, Lowes Companies, Inc., Mooresville, NC
Sandra Rosenbloom, Professor, University of Texas, Austin
Henry G. (Gerry) Schwartz, Jr., Chairman (retired), Jacobs/Sverdrup Civil, Inc., St. Louis, MO
Kumares C. Sinha, Olson Distinguished Professor of Civil Engineering, Purdue University, West Lafayette, IN
Gary C. Thomas, President and Executive Director, Dallas Area Rapid Transit, Dallas, TX
Paul Trombino III, Director, Iowa DOT, Ames
Phillip A. Washington, General Manager, Regional Transportation District, Denver, CO

EX OFFICIO MEMBERS
Thomas P. Bostick (Lt. General, U.S. Army), Chief of Engineers and Commanding General, U.S. Army Corps of Engineers, Washington, DC
Alison Jane Conway, Assistant Professor, Department of Civil Engineering, City College of New York, NY, and Chair, TRB Young Member Council
Anne S. Ferro, Administrator, Federal Motor Carrier Safety Administration, U.S. DOT
David J. Friedman, Acting Administrator, National Highway Traffic Safety Administration, U.S. DOT
LeRoy Gishi, Chief, Division of Transportation, Bureau of Indian Affairs, U.S. Department of the Interior
John T. Gray II, Senior Vice President, Policy and Economics, Association of American Railroads, Washington, DC
Michael P. Huerta, Administrator, Federal Aviation Administration, U.S. DOT
Paul N. Jaenichen, Sr., Acting Administrator, Maritime Administration, U.S. DOT
Therese W. McMillan, Acting Administrator, Federal Transit Administration, U.S. DOT
Michael P. Melaniphy, President and CEO, American Public Transportation Association, Washington, DC
Gregory G. Nadeau, Acting Administrator, Federal Highway Administration, U.S. DOT
Cynthia L. Quarterman, Administrator, Pipeline and Hazardous Materials Safety Administration, U.S. DOT
Peter M. Rogoff, Under Secretary for Policy, U.S. DOT
Craig A. Rutland, U.S. Air Force Pavement Engineer, Air Force Civil Engineer Center, Tyndall Air Force Base, FL
Joseph C. Szabo, Administrator, Federal Railroad Administration, U.S. DOT
Barry R. Wallerstein, Executive Officer, South Coast Air Quality Management District, Diamond Bar, CA
Gregory D. Winfree, Assistant Secretary for Research and Technology, Office of the Secretary, U.S. DOT
Frederick G. (Bud) Wright, Executive Director, American Association of State Highway and Transportation Officials, Washington, DC
Paul F. Zukunft (Adm., U.S. Coast Guard), Commandant, U.S. Coast Guard, U.S. Department of Homeland Security

* Membership as of August 2014.

N AT I O N A L C O O P E R AT I V E H I G H W AY R E S E A R C H P R O G R A M

NCHRP REPORT 782


Proposed Guideline
for Reliability-Based
Bridge Inspection Practices
Glenn Washer
Massoud Nasrollahi
Christopher Applebury
University of Missouri
Columbia, MO

Robert Connor
Purdue University
West Lafayette, IN

Adrian Ciolko
KPFF Consulting Engineers
Evanston, IL

Robert Kogler
Rampart, LLC
Arlington, VA

Philip Fish
Fish and Associates, Inc.
Middleton, WI

David Forsyth
TRI/Austin
Austin, TX

Subscriber Categories

Bridges and Other Structures

Research sponsored by the American Association of State Highway and Transportation Officials
in cooperation with the Federal Highway Administration

TRANSPORTATION RESEARCH BOARD


WASHINGTON, D.C.
2014
www.TRB.org

NATIONAL COOPERATIVE HIGHWAY


RESEARCH PROGRAM

NCHRPREPORT 782

Systematic, well-designed research provides the most effective


approach to the solution of many problems facing highway
administrators and engineers. Often, highway problems are of local
interest and can best be studied by highway departments individually
or in cooperation with their state universities and others. However, the
accelerating growth of highway transportation develops increasingly
complex problems of wide interest to highway authorities. These
problems are best studied through a coordinated program of
cooperative research.
In recognition of these needs, the highway administrators of the
American Association of State Highway and Transportation Officials
initiated in 1962 an objective national highway research program
employing modern scientific techniques. This program is supported on
a continuing basis by funds from participating member states of the
Association and it receives the full cooperation and support of the
Federal Highway Administration, United States Department of
Transportation.
The Transportation Research Board of the National Academies was
requested by the Association to administer the research program
because of the Boards recognized objectivity and understanding of
modern research practices. The Board is uniquely suited for this
purpose as it maintains an extensive committee structure from which
authorities on any highway transportation subject may be drawn; it
possesses avenues of communications and cooperation with federal,
state and local governmental agencies, universities, and industry; its
relationship to the National Research Council is an insurance of
objectivity; it maintains a full-time research correlation staff of specialists
in highway transportation matters to bring the findings of research
directly to those who are in a position to use them.
The program is developed on the basis of research needs identified
by chief administrators of the highway and transportation departments
and by committees of AASHTO. Each year, specific areas of research
needs to be included in the program are proposed to the National
Research Council and the Board by the American Association of State
Highway and Transportation Officials. Research projects to fulfill these
needs are defined by the Board, and qualified research agencies are
selected from those that have submitted proposals. Administration and
surveillance of research contracts are the responsibilities of the National
Research Council and the Transportation Research Board.
The needs for highway research are many, and the National
Cooperative Highway Research Program can make significant
contributions to the solution of highway transportation problems of
mutual concern to many responsible groups. The program, however, is
intended to complement rather than to substitute for or duplicate other
highway research programs.

Project 12-82(01)
ISSN 0077-5614
ISBN 978-0-309-30791-8
Library of Congress Control Number 2014946785
2014 National Academy of Sciences. All rights reserved.

COPYRIGHT INFORMATION
Authors herein are responsible for the authenticity of their materials and for obtaining
written permissions from publishers or persons who own the copyright to any previously
published or copyrighted material used herein.
Cooperative Research Programs (CRP) grants permission to reproduce material in this
publication for classroom and not-for-profit purposes. Permission is given with the
understanding that none of the material will be used to imply TRB, AASHTO, FAA, FHWA,
FMCSA, FTA, or Transit Development Corporation endorsement of a particular product,
method, or practice. It is expected that those reproducing the material in this document for
educational and not-for-profit uses will give appropriate acknowledgment of the source of
any reprinted or reproduced material. For other uses of the material, request permission
from CRP.

NOTICE
The project that is the subject of this report was a part of the National Cooperative Highway
Research Program, conducted by the Transportation Research Board with the approval of
the Governing Board of the National Research Council.
The members of the technical panel selected to monitor this project and to review this
report were chosen for their special competencies and with regard for appropriate balance.
The report was reviewed by the technical panel and accepted for publication according to
procedures established and overseen by the Transportation Research Board and approved
by the Governing Board of the National Research Council.
The opinions and conclusions expressed or implied in this report are those of the
researchers who performed the research and are not necessarily those of the Transportation
Research Board, the National Research Council, or the program sponsors.
The Transportation Research Board of the National Academies, the National Research
Council, and the sponsors of the National Cooperative Highway Research Program do not
endorse products or manufacturers. Trade or manufacturers names appear herein solely
because they are considered essential to the object of the report.

Published reports of the

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM


are available from:
Transportation Research Board
Business Office
500 Fifth Street, NW
Washington, DC 20001
and can be ordered through the Internet at:
http://www.national-academies.org/trb/bookstore
Printed in the United States of America

The National Academy of Sciences is a private, nonprofit, self-perpetuating society of distinguished scholars engaged in scientific
and engineering research, dedicated to the furtherance of science and technology and to their use for the general welfare. Upon the
authority of the charter granted to it by the Congress in 1863, the Academy has a mandate that requires it to advise the federal
government on scientific and technical matters. Dr. Ralph J. Cicerone is president of the National Academy of Sciences.
The National Academy of Engineering was established in 1964, under the charter of the National Academy of Sciences, as a parallel
organization of outstanding engineers. It is autonomous in its administration and in the selection of its members, sharing with the
National Academy of Sciences the responsibility for advising the federal government. The National Academy of Engineering also
sponsors engineering programs aimed at meeting national needs, encourages education and research, and recognizes the superior
achievements of engineers. Dr. C. D. Mote, Jr., is president of the National Academy of Engineering.
The Institute of Medicine was established in 1970 by the National Academy of Sciences to secure the services of eminent members
of appropriate professions in the examination of policy matters pertaining to the health of the public. The Institute acts under the
responsibility given to the National Academy of Sciences by its congressional charter to be an adviser to the federal government
and, upon its own initiative, to identify issues of medical care, research, and education. Dr. Victor J. Dzau is president of the
Institute of Medicine.
The National Research Council was organized by the National Academy of Sciences in 1916 to associate the broad community of
science and technology with the Academys purposes of furthering knowledge and advising the federal government. Functioning in
accordance with general policies determined by the Academy, the Council has become the principal operating agency of both the
National Academy of Sciences and the National Academy of Engineering in providing services to the government, the public, and
the scientific and engineering communities. The Council is administered jointly by both Academies and the Institute of Medicine.
Dr. Ralph J. Cicerone and Dr. C. D. Mote, Jr., are chair and vice chair, respectively, of the National Research Council.
The Transportation Research Board is one of six major divisions of the National Research Council. The mission of the Transportation Research Board is to provide leadership in transportation innovation and progress through research and information exchange,
conducted within a setting that is objective, interdisciplinary, and multimodal. The Boards varied activities annually engage about
7,000 engineers, scientists, and other transportation researchers and practitioners from the public and private sectors and academia,
all of whom contribute their expertise in the public interest. The program is supported by state transportation departments, federal
agencies including the component administrations of the U.S. Department of Transportation, and other organizations and individuals interested in the development of transportation. www.TRB.org

www.national-academies.org

COOPERATIVE RESEARCH PROGRAMS

CRP STAFF FOR NCHRP REPORT 782


Christopher W. Jenks, Director, Cooperative Research Programs
Christopher Hedges, Manager, National Cooperative Highway Research Program
Waseem Dekelbab, Senior Program Officer
Danna Powell, Senior Program Assistant
Eileen P. Delaney, Director of Publications
Scott E. Hitchcock, Editor

NCHRP PROJECT 12-82(01) PANEL


Field of DesignArea of Bridges
Matthew Farrar, Idaho Transportation Department, Boise, ID (Chair)
Laura M. Amundson, Parsons Brinckerhoff, Minneapolis, MN
William R. Randy Cox, American Segmental Bridge Institute, Driftwood, TX
Arthur W. DAndrea, Louisiana DOTD, Baton Rouge, LA
David A. Juntunen, Michigan DOT, Lansing, MI
Peter C. McCowan, New York State DOT, Albany, NY
Barton J. Newton, California DOT, Sacramento, CA
Andre V. Pavlov, Florida DOT, Tallahassee, FL
Lance D. Savant, AECOM, Mechanicsburg, PA
Thomas D. Everett, FHWA Liaison
James W. Bryant, Jr., TRB Liaison

AUTHOR ACKNOWLEDGMENTS
The research reported herein was performed under NCHRP Project 12-82 by the Department of Civil
and Environmental Engineering at the University of Missouri (MU). The University of Missouri was the
Contractor for this study. Dr. Glenn Washer, Associate Professor of Civil Engineering at MU, was the
Principal Investigator. Dr. Robert Connor, Associate Professor of Civil Engineering at Purdue University,
was the Co-Principal Investigator. The other authors of this report are Adrian Ciolko, Group Manager
Bridges & Infrastructure at KPFF Consulting Engineers; Robert Kogler, Principal at Rampart, LLC.; Philip
Fish, Founder and Chairman at Fish and Associates, Inc.; David Forsyth, NDE Division Manager at TRI/
Austin; Christopher Applebury, Research Assistant and M.S. Candidate at MU; and Massoud Nasrollahi,
Research Assistant and Ph.D. Candidate at MU.
The Research Team gratefully acknowledges the helpful insights and comments provided by the project
panel during the course of the research. The Research Team would also like to acknowledge the assistance
provided by the Texas and Oregon Departments of Transportation during the execution of the case study
portions of the research.

FOREWORD

By Waseem Dekelbab
Staff Officer
Transportation Research Board

This report presents a proposed Guideline for reliability-based bridge inspection practices and provides two case studies of the application of the proposed Guideline. The Guideline describes a methodology to develop a risk-based approach for determining the bridge
inspection interval according to the requirements in the Moving Ahead for Progress in
the 21st Century Act (MAP-21) legislation. The goal of the methodology is to improve
the safety and reliability of bridges by focusing inspection efforts where most needed and
optimizing the use of resources. The material in this report will be of immediate interest to
bridge engineers.
The National Bridge Inspection Standards (NBIS) mandate the frequency and methods
used for the safety inspection of highway bridges. The inspection intervals specified in the
NBIS require routine inspections to be conducted every 24 months, and that interval may
be extended to 4 years for bridges that meet certain criteria and are approved by FHWA. For
bridges with fracture-critical elements, hands-on inspections are required every 2 years. The
specified intervals are generally not based on performance of bridge materials or designs, but
rather on experience from managing almost 600,000 bridges in the National Bridge Inventory.
These inspection intervals are applied to the entire bridge inventory, but they may not be
appropriate for all bridges. For example, recently constructed bridges typically experience
few problems during their first decade of service and those problems are typically minor.
Under the present requirements, these bridges must have the same inspection frequency
and intensity as a 50-year-old bridge that is reaching the end of its service life. In the case of
bridges with fracture-critical elements, newer bridges with improved fabrication processes
and designs that mitigate the effects of fatigue are inspected on the same interval and to the
same intensity as older bridges that do not share these characteristics.
A more rational approach to determining appropriate inspection practices for bridges
would consider the structure type, age, condition, importance, environment, loading,
prior problems, and other characteristics of the bridge. There is a growing consensus that
these inspection practices should meet two goals: (1) improving the safety and reliability
of bridges and (2) optimizing resources for bridge inspection. These goals can be accomplished through the application of reliability theory.
Research was performed under NCHRP Projects 12-82 and 12-82(01) by the University
of Missouri to develop a proposed bridge inspection practice for consideration for adoption
by AASHTO. The methodology developed is based on rational methods to ensure bridge
safety, serviceability, and effective use of resources.
The report includes two parts: Part IProposed Guideline for Reliability-Based Bridge
Inspection Practices and Part IIFinal Research Report: Developing Reliability-Based
Inspection Practices that documents the entire research effort.

CONTENTS

P A R T I Proposed Guideline for Reliability-Based Bridge


Inspection Practices
3 Summary
4 Definitions
6





7
7
8
8
10
11

12



12
13
14
15

17
17
18
19
19
19
19
20

21





21
21
23
23
23
24

Chapter 1Introduction
1.1Process
1.1.1Scope
1.1.2Purpose
1.2Background
1.2.1Reliability and Probability
1.2.2Consequences

Chapter 2Reliability Assessment of Bridge Elements


2.1Definition of Failure
2.2Occurrence Factors
2.3Assessment of Consequences
2.4The Reliability Assessment Panel

Chapter 3Determination of Inspection Interval and Scope


3.1Inspection Interval
3.1.1Inspection Scope
3.1.2Sampling
3.1.3Maintenance Inspections
3.1.4Initial Inspections
3.1.5Start-Up Inspections
3.1.6Quality Control/Quality Assurance

Chapter 4Establishing an RBI Program


4.1Overview of Process
4.2Setting the Scope of the Analysis
4.3Training Requirements
4.3.1Training for RAP Members
4.3.2Training of Inspectors
4.4Software Development and Integration

25 References
26

Appendix AGuideline for Evaluating the Occurrence Factor

39

Appendix BGuideline for Evaluating the Consequence Factor

51

Appendix CGuideline for Determining the Inspection Interval

54

Appendix DInspection Technologies

59

Appendix EAttribute Index and Commentary

97

Appendix FIllustrative Examples

P A R T I I Final Research Report: Developing


Reliability-Based Inspection Practices
131 Summary
135

Chapter 1Background

137

Chapter 2Research Approach

139

Chapter 3Findings and Applications

139
140
141
142
143
144
144
146
146
150
151
153
159
159
160
161
164
164
164
164
167
169
170
170
173
175
176
176
176
177
182
183
183
184
188
188

3.1Introduction
3.2Overview of Methodology
3.3Reliability
3.3.1Reliability Theory
3.3.2Failure
3.3.3Damage Modes and Deterioration Mechanisms
3.3.4Lifetime Performance Characteristics
3.4Key Elements of RBI
3.4.1The OF
3.4.2CFs
3.4.3Inspection Procedures in RBI
3.4.4RAP
3.5Data to Support RBI Analysis
3.5.1Quantitative vs. Qualitative Analysis
3.5.2Data Needed for Assessment
3.5.3Industry Data
3.6Case Studies of the Methodology
3.6.1Summary Overview of RAP Meeting
3.6.2RAP Meeting Attendees
3.6.3Schedule and Agenda
3.6.4Back-Casting Procedure
3.6.5Statistical Analysis of NBI Data
3.6.6Bridge Inventories in Texas and Oregon
3.6.7Time-in-Condition Rating
3.6.8Overview of Case Study Results
3.6.9CFs
3.7Back-Casting Results for Oregon
3.7.1Environments
3.7.2CFs
3.7.3Back-Casting Results for Oregon
3.8Back-Casting Results for Texas
3.8.1Environments
3.8.2CFs
3.8.3Back-Casting Results for Texas
3.9Discussion of the Case Studies in Texas and Oregon
3.9.1Back-Casting Results

189
189
190
190

Chapter 4Conclusions, Recommendations,


and Suggested Research
4.1Recommendations
4.2Suggested Research
4.2.1Implementation Strategy

195 References
197 Abbreviations
198

Appendix ADeveloping Reliability-Based Inspection Practices:


Oregon Pre-Stressed Bridges

204

Appendix BTexas Steel Bridge Attributes Summary

210

Appendix CControlling Damage Modes for Sample Bridges

Note: Many of the photographs, figures, and tables in this report have been converted from color to grayscale
for printing. The electronic version of the report (posted on the Web at www.trb.org) retains the color versions.

Pa rt I

Proposed Guideline for


Reliability-Based Bridge
Inspection Practices

SUMMARY

Proposed Guideline for


Reliability-Based Bridge
Inspection Practices
This guideline describes a methodology for developing Risk-Based Inspection (RBI) practices for highway bridges. The goal of the methodology is to improve the safety and reliability of bridges by focusing inspection efforts where most needed and optimizing the use of
resources. The guideline provides the framework and procedures for conducting reliability
assessments to develop suitable inspection strategies for bridges based on an engineering
assessment of inspection needs. The methodology considers the structure type, age, condition, importance, environment, loading, prior problems, and other characteristics that
contribute to the reliability and durability of highway bridges.
RBI practices differ from traditional approaches that are generally calendar based, because
the setting of inspection frequencies (or intervals) and scope are not fixed or uniform.
Rather, reliability-based engineering analysis is conducted to assess the inspection needs
for a particular bridge or family of bridges, and inspection requirements, i.e., frequency
and scope, are aligned with those needs. This is achieved by analyzing the likelihood of
anticipated damage modes and the associated outcomes or consequences. As a result, RBI
practices can focus attention specifically on the damage and deterioration mechanisms that
are most important for ensuring bridge safety. As such, they provide a better linkage between
damage modes that affect bridges and the inspection approaches that will best reduce the
associated risks, leading to improved bridge safety. This approach has been widely accepted
in many industries with facilities that can be considered analogous to highway bridges: very
large, expensive, and complex structural systems that are exposed to rugged environmental
conditions and mechanical loading.
The purpose of this document is to provide guidance for bridge owners for conducting reliability-based assessments to determine inspection needs. The methodology requires
bridge owners to perform a reliability assessment of bridges within their bridge inventory to
identify those bridges that are most in need of inspection to ensure bridge safety, and those
where inspection needs are less. An expert panel assembled at the owner level performs
this assessment. The assessment considers the reliability and safety attributes of the bridges
to assess the likelihood of damage and evaluate the potential outcomes or consequences in
terms of safety and serviceability. Through this process, inspection needs are prioritized to
improve the safety and reliability of the bridge inventory overall.

Definitions

Attributes: Characteristics that affect the reliability of a bridge or bridge element.


Condition Attributes: Characteristics that relate to the current condition of a bridge or bridge
element. These may include element ratings, component ratings, and specific damage modes or
mechanisms that have a significant effect on the reliability of an element.
Consequence Factor: Factor describing the expected outcome or result of a failure.
Damage Mode: Typical damage affecting the condition of a bridge element (e.g., spalling of
concrete, cracking, etc.).
Design Attributes: Characteristics of bridge or bridge element that are part of the elements
design. These attributes typically do not change over time except when renovation, rehabilitation,
or preservation activities occur.
Deterioration Mechanism: Process or phenomena resulting in damage to a bridge element
(e.g., corrosion, fatigue, etc).
Element: Identifiable portions of a bridge made of the same material, having similar role in the
performance of the bridge, and expected to deteriorate in a similar fashion.
Failure: Termination of the ability of a system, structure, or component to perform its intended
function (1). For bridges, the condition at which a given bridge element is no longer performing
its intended function to safely and reliably carry normal loads and maintain serviceability.
Loading Attributes: Loading characteristics that affect the reliability of a bridge or bridge element, such as traffic or environment.
Occurrence Factor: Factor describing the likelihood that an element will fail during a specified
time period.
Operational Environment: The operational environment is a combination of the circumstances
surrounding and potentially affecting the in-service performance of bridges and bridge elements.
These include typical loading patterns, ambient environmental conditions, construction quality
and practices, maintenance and management practices, and other factors that may vary between
different geographic regions and/or organizational boundaries.
Probability: Extent to which an event is likely to occur during a given time interval (1). This
may be based on the frequency of events, or on degree of belief or expectation. Degrees of belief
about probability can be chosen using qualitative scales, ranks, or categories such as remote/
low/moderate/high.
Reliability: Ability of an element or component to operate safely under designated operating
conditions for a designated period of time.
4

Definitions5

Risk: Combination of the probability of an event and its consequence.


Risk Analysis: Systematic use of information to estimate the risk. Sources of information may
include historical data, theoretical analysis and engineering judgment.
Screening Attribute: Characteristics of a bridge or bridge element that:
Make the likelihood of serious damage unusually high,
Make the likelihood of serious damage unusually uncertain, and
Identify a bridge with different anticipated deterioration patterns than other bridges in a

group or family.

CHAPTER 1

Introduction

This guideline describes a methodology for developing Risk-Based Inspection (RBI) practices for highway bridges. The goal of the methodology is to improve the safety and reliability of bridges by focusing inspection efforts where most needed and optimizing the use
of resources. The guideline provides a framework and procedures for developing suitable
inspection strategies based on a rational engineering assessment of inspection needs. The
methodology considers the structure type, age, condition, importance, environment, loading,
prior problems, and other characteristics that contribute to the reliability and durability of
highway bridges.
The methodology requires bridge owners to perform a reliability assessment of bridges within
their bridge inventory to identify those bridges that are most in need of inspection to ensure
bridge safety, and those where inspection needs are less. This assessment is conducted by considering the reliability and safety attributes of bridges, assessing the likelihood of damage and
associated deterioration mechanisms, and evaluating the potential outcomes or consequences
in terms of safety and serviceability. Through this process, inspection needs are prioritized to
improve the safety and reliability of the bridge inventory overall.
This chapter of the document provides an introduction and overview of the process, as well
as background information on the underlying theories and common practices for RBI and reliability assessments. Chapter 2 of the document describes the methodology for conducting a reliability assessment for bridges. This includes providing a definition of element failure suitable as
an analysis tool, and a description of the key factors to be assessed in the typical reliability assessment conducted for inspection planning purposes. This chapter also describes the composition
of the Reliability Assessment Panel (RAP) that will conduct the assessments.
Chapter 3 describes the process for determining the appropriate maximum inspection interval
and scope of inspection, based on analysis as described in Chapter 2. The underlying approaches
for identifying inspection intervals and the techniques or methods to be used for the inspections are discussed. Finally, Chapter 4 provides an overview of the overall process, guidance for
bridge owners on beginning an RBI program, transitioning from traditional, calendar-based
approaches, and general guidance on the training that may be required.
There are six appendices in the document that describe in more detail the process and
mechanics of the analysis. Guidance for determining the factors necessary to perform a reliability assessment are included in Appendices A, B, and C. Guidance on inspection methods
and nondestructive evaluation (NDE) technologies that can be used for conducting RBIs is
described in Appendix D. Appendix E contains commentary regarding specific, common
attributes of bridges that influence damage modes and deterioration mechanisms, and relate
to bridge reliability. Finally, Appendix F includes three example implementations of the
methodology applied to bridges of common design: a multi-girder concrete bridge with
6

Introduction7

prestressed superstructure elements constructed in the past 5 years, a multi-girder steel


bridge constructed more than 50 years ago, and a multi-girder reinforced concrete bridge
constructed in 1963.

1.1Process
The process involves an owner (e.g., state) establishing an expert panel to define and assess
the durability and reliability characteristics of bridges within the state. The expert group analyzes
portions of the bridge inventory to assess inspection needs by using engineering rationale, experience, and typical deterioration patterns to evaluate the reliability characteristics of bridges and
the potential outcomes of damage. This is done through a relatively simple process that consists
of three primary steps:
Step 1: What can go wrong, and how likely is it? Identify possible damage modes for the elements of a selected bridge type. Considering design, loading, and condition characteristics
(attributes), categorize the likelihood of serious damage occurring into one of four occurrence
factors (OFs) ranging from remote (very unlikely) to high (very likely).
Step 2: What are the consequences? Assess the consequences, in terms of safety and serviceability, assuming the given damage modes occur. Categorize the potential consequences into one
of four consequence factors (CFs) ranging from low (minor effect on serviceability) through
severe (i.e., bridge collapse, loss of life).
Step 3: Determine the inspection interval and scope. Use a simple 4 4 matrix to prioritize
inspection needs and assign an inspection interval for the bridge based on the results of
Steps 1 and 2. Damage modes that are likely to occur and have high consequences are prioritized over damage modes that are unlikely to occur or are of little consequence in terms of
safety. An RBI procedure is developed based on the assessment of typical damage modes for
the bridges being assessed.
Inspections are conducted according to the RBI procedure developed through this process.
Results of the inspection are assessed to determine if the existing RBI procedure needs to be
modified or updated as a result of findings from the inspection.
Through this process, individual bridges, or groups of bridges of similar design characteristics,
can be assessed to evaluate the inspection needs based on an engineering analysis of the likelihood
of serious damage occurring and the effect of that damage on the safety and serviceability of the
bridge. This approach considers the structure type, age, condition, and operational environment
in a systematic manner to provide a rational assessment process for inspection planning. A documented rationale for the inspection strategy utilized for a given bridge is developed. The damage
modes most important to ensuring the safety and serviceability of the bridge are identified such
that inspection efforts can be focused to improve the reliability of the inspection results.

1.1.1Scope
This guide is focused on the inspection of typical highway bridges of common design characteristics. Atypical structures, such as long-span truss bridges, cable-stayed bridges, suspension bridges, and other unique or unusual bridge designs may require certain considerations
not presently captured in this guideline; this guideline provides for inspection planning for
the superstructure, substructure, and deck for typical highway bridges. Scour and underwater
inspections have existing methodologies for evaluation, and, as such, are not included herein.
Bridges assessed using this methodology are assumed to have a current load rating that indicates
that the structural capacity is sufficient to carry allowable loads.

8 Proposed Guideline for Reliability-Based Bridge Inspection Practices

1.1.2Purpose
The purpose of this document is to provide guidance for bridge owners for conducting
reliability-based assessments for determining the frequency and scope of inspections for typical
highway bridges. This document is intended to be used by bridge owners for assessing their
bridge inventories in order to prioritize inspection needs based on an engineering analysis that
considers the bridge type, age, loading, condition, and other characteristics of a bridge. This
guideline is intended for application to typical bridges with common and ordinary forms of
deterioration and damage. Advanced deterioration and/or specific defects such as fatigue cracks
due to primary stresses or severe corrosion damage in concrete typically require more detailed
engineering analysis than provided herein.

1.2Background
The periodic inspection of highway bridges in the United States plays a critical role in ensuring
the safety, serviceability, and reliability of bridges. Inspection processes have developed over time
to meet the requirements of the National Bridge Inspections Standards (NBIS)(2) and to meet the
needs of individual bridge owners in terms of managing and maintaining their bridge inventory.
The inspection frequency mandated by the NBIS requires the inspection interval (maximum time
period between inspections) not to exceed 24 months. Based on certain criteria, that interval may
be extended up to 48 months with approval from the Federal Highway Administration (FHWA)
(3). Maximum inspection intervals of less than 24 months are utilized for certain bridges according to criteria developed by the bridge owner, typically based on age and known deficiencies. Most
bridge owners utilize the uniform maximum inspection interval of 24 months, as mandated by
the NBIS, for the majority of the bridges in their inventory, and the reduced intervals for bridges
with known deficiencies. Only 15 states utilize the 48 month policy, often only for culverts. The
uniform inspection interval of 24 months was specified at the origination of the National Bridge
Inspection Program in 1971 based on experience, engineering judgment, and the best information available at the time. The uniform approach provides a single maximum inspection interval
for most bridges, regardless of the bridge age, design, or environment. To date, this mandated
inspection interval has provided an adequate level of safety and reliability for the bridge inventory nationwide. However, such a uniform inspection interval does not consider explicitly the
likelihood of failure based on bridge condition, design, or operating environment, or the potential consequences of a failure. A uniform inspection interval does not recognize that a newly
constructed bridge with improved durability characteristics and a few years of exposure to the
service environment may be much less likely to develop serious damage over a given time interval
than an older bridge that has been exposed to the service environment for many years. Bridges
that are in benign, arid operating environments are inspected at the same interval as bridges in
aggressive marine environments, where significant damage from corrosion may develop much
more rapidly. Current practices make it difficult to distinguish if the same or improved safety
and reliability could be achieved by varying inspection methods or frequencies to meet the needs
of a specific bridge based on its design and operational environment. The current approach also
makes it difficult to analyze if a given inspection activity is excessive, or if it provides little or no
measure of increased assurance of the safety and reliability of bridges. Given that any inspection
activity carries with it a certain amount of risk to both the inspector and to the traveling public,
inspections that are excessive or that provide little benefit may present added, unnecessary risks.
Otherwise, inspections that are inadequate or fail to distinguish the importance of critical damage
modes may also present certain added risks that require analysis.
Recognizing the variability in the design, condition, and operating environments of bridges
would provide for inspection requirements that better meet the needs of individual bridges to

Introduction9

improve both bridge and inspection reliability. Other industries are increasingly recognizing
the limitations of prescribed inspection frequencies and are developing methodologies for efficiently assessing inspection needs, ensuring the safety and reliability of systems, and focusing
inspection resources most effectively (1, 46). Methodologies for assessing inspection needs
based on the likelihood of a service failure, combined with the consequences of such a failure,
is a common approach to inspection planning and to developing effective inspection strategies.
These approaches are typically described as risk-based, where inspection planning is conducted
considering the reliability of a component, i.e., how likely is it that the component or machine
will fail during a certain time period, and the consequences of such an event. Damage modes and
deterioration mechanisms are typically assessed explicitly to determine the likelihood of failure
during a given time period, and to identify the appropriate inspection methods to detect critical
damage prior to failure.
A risk-based approach has been adopted in many industries as a tool for inspection planning, to focus attention on the component or machine that represents the greatest risk. Risk is
defined as the product of the probability of an event and the associated consequences:
Risk = Probability Consequence
Probability in this equation is the likelihood of an adverse event or failure occurring during a
given time period. This is sometimes expressed quantitatively as a probability of failure (POF)
estimate for a given time interval, or as a qualitative assessment of the likelihood of an adverse
event based on experience and engineering judgment. Consequence is a measure of the impact
of the event occurring, which may be measured in terms of economic, social, safety, or environmental impacts.
Risk can be expressed quantitatively using POF estimates or models and quantitative measures of consequences, such as the cost of a certain event or the loss of service of a component.
Risk can also be expressed qualitatively by estimating whether the likelihood of a certain event
is high, medium, or low, and determining a qualitative estimate of the consequences. Presenting risk qualitatively is a common and effective method for evaluating risk and for assessing
relative risk efficiently. Figure 1 shows a qualitative risk matrix (1, 5). This matrix shows a good
representation of the overall concept and basic principles of risk. A high likelihood (probability)
of occurrence combined with a high consequence results in a high risk, located in the upper
right corner of the figure. Low likelihood combined with a low consequence results in low risk,
located in the lower left-hand corner of the figure. High risk and low risk elements typically
do not create challenges in decision making; items that are high risk may not be acceptable
and actions are required to lower the risk, either by reducing the likelihood of an event, or by
reducing the consequences, or both. Items that are low risk are typically acceptable and may
require little or no action. In the medium risk area, questions may arise about how much risk
is acceptable, and what the appropriate decision-making strategies are for mitigating that risk.
In terms of inspection strategies, items that are high risk are prioritized for more frequent and
possibly more intense inspections to reduce uncertainty and to monitor the development of
damage to ensure that safety is maintained. Items that are low risk may have longer inspection
intervals and have less intense inspection protocols.
An important concept in risk analysis is to understand that high likelihood does not necessarily mean high risk, if the consequences are small. Similarly, high consequence does not necessarily mean high risk, if the likelihood is small. The level of risk can only be determined once both
of these variables are assessed.
A risk-based planning approach focuses attention not on the items that are most likely to fail,
but rather those items whose failure is most important, by considering both the likelihood of

Figure 1.Risk
matrix showing
high-, medium-,
and low-risk values.

10 Proposed Guideline for Reliability-Based Bridge Inspection Practices

failure and the associated consequences. The setting of inspection frequencies or intervals is not
a rigid process, such as is the case for uniform or calendar-based inspection frequencies. Rather,
it is a process that evolves and changes over the life of a component such that inspection frequencies change as risk increases (or decreases). Therefore, the frequency of inspection is aligned with
the needs and the associated risks, focusing attention on the most at-risk items. This approach
has been widely accepted in many industries with facilities that can be considered analogous to
highway bridges: very large, expensive, and complex structural systems that are exposed to rugged environmental conditions and mechanical loading (1, 4, 6).

1.2.1 Reliability and Probability


Reliability is defined as the ability of an item to operate safely under designated operating
conditions for a designated period of time or number of cycles. For bridges and bridge elements,
reliability typically decreases as a function of time due to deterioration and the damage accumulated during the service life of a bridge. That is, the likelihood of failure typically increases with
time as the element ages, due to deterioration mechanisms such as corrosion or fatigue. The
reliability of a bridge or bridge element can be expressed as:
R ( t ) = Pr (T t )
Where R(t) is the reliability, T is the time to failure for the item, and t is the designated period
of time for the items operation. In other words, the reliability is the probability (Pr) or likelihood that the failure time exceeds the operation time. Sometimes, the probability is expressed
as a probability density function (pdf) that expresses the time to failure of an item (T) as some
generic distribution, such as normal, log normal, etc. This distribution can be used to calculate a
POF function, F(t), to express the probability that the item will fail sometime up to time t. This
time-varying function describes likelihood of failure up to some given time, or the unreliability
of the item, and the reliability is then:
R (t ) = 1 F (t )
In other words, the reliability is the probability that the item will not fail during the time
period of interest. When a large population of test data of identical or near identical components
exposed to the same operational environment are available, a probability function describing the
failure characteristics of the component may be determined and verified based on the results. If
test data are not available, a suitable distribution must be assumed based on the general characteristics of the population, typical failure behavior, and known deterioration mechanisms.
These distributions are typically based on experience and assumptions regarding the anticipated
performance of the system or component. This is challenging and can lead to unsubstantiated
confidence in the model when the design characteristics, construction quality, condition, and
operational environment of the components vary. Even if substantial data were readily available, design and construction practices are constantly evolving such that past performance may
not indicate future performance. Critical damage modes may have yet to manifest in observable
damage, and as such may not be included in the data. Given the large variation in the design,
construction, construction quality, and operational environments, the utility of probabilistic
models to effectively predict the future performance of a specific bridge or bridge element is
problematic.
Under these circumstances, engineering judgment and experience is needed to estimate the
expected reliability of a specific component, or set of components, of similar design and construction quality operating within a specific operational environment. Engineering judgment

Introduction11

is required to estimate the reliability of bridge elements based on past experience, engineering
knowledge, and a rational process to systematically assess bridges of common design and construction characteristics. The process involves engineers with experience and expertise in the
performance of bridges within a particular operational environment using engineering judgment to assess the probability (likelihood) of failure during some future time period. When
combined with an assessment of the consequences, an effective analysis can be conducted to
identify inspection needs efficiently.

1.2.2Consequences
The primary purpose of bridge inspection is to ensure the safety and serviceability of highway
bridges. As a result, the consequences to be assessed in prioritizing the importance of different
damage modes are assessed in terms of bridge safety and serviceability. The consequence of failure, or of serious damage developing in a bridge element, typically depends on the role of that
element in the structural system of the bridge, and on the operating environment surrounding
the bridge. For example, the consequence of an abutment having severe corrosion damage might
be low, while the same damage in a main superstructure member may be high. The consequence
of damage developing at the soffit of a bridge deck, such as concrete spalling, might be low if the
bridge is over a flood plain, but high if the bridge is over an interstate highway. The consequence
associated with a given damage mode can be assessed through engineering judgment, through
common or related experience, or through theoretical analysis.
The process developed and described herein requires the determination of two key parameters: an estimate of the reliability of given bridge elements, based on the likelihood (probability)
that the element would fail during a given time interval, and an assessment of the consequences of
that failure. These data are then used to determine an appropriate inspection interval and scope
(procedures and methods) for a bridge. As such, the methodology described is a reliability-based
bridge inspection planning process for ensuring the safety and serviceability (i.e., reliability) of
highway bridges.

CHAPTER 2

Reliability Assessment
of Bridge Elements
This section describes the methodology for reliability assessment of the bridge elements. Section 2.1 describes and defines failure as applied to typical bridge elements for the reliability
assessment. Section 2.2 describes the methodology for evaluating the probability or likelihood
that failure will occur (OF). Section 2.3 describes the methodology for evaluating the consequences of that occurrence (CFs). Finally, Section 2.4 discusses the panel that conducts the
assessment, the RAP.

2.1 Definition of Failure


It is critical that the conditions that constitute a failure be defined before beginning a reliability assessment. For bridges, catastrophic collapse would be one obvious condition that could be
used to define failure. For most bridges, the probability of catastrophic failure is very remote.
For bridge inspections, important concerns extend well beyond simply avoiding catastrophic
failure. Ensuring the safety of the bridge, the safety of those traveling on or below the bridge, and
the serviceability of the bridge are each critical. Maintenance and repair activities are needed to
support the serviceability of the bridge and ensure the safety of motorists, even while the likelihood of a catastrophic failure remains remote.
Therefore, failure requires a suitable definition that captures the need to ensure the structural safety of the bridge, the safety of travelers on or below the bridge, and the serviceability
of the bridge. Failure, utilized in this context, is defined as when an element is no longer performing its intended function to safely and reliably carry normal loads and maintain serviceability.
For example, a bridge deck with severe spalling may represent a failed condition for the bridge
deck even though the deck may have adequate load-carrying capacity, because the ability of
the deck to reliably carry traffic is compromised. The condition rating of 3, serious condition according to the NBIS rating system, is used in the analysis described herein as a general
description of a failed condition. It is not envisioned that any bridges or bridge elements
assessed using a risk-based approach are allowed to deteriorate to this condition. Rather, inspection intervals are adjusted to ensure that the likelihood of failure in the time intervals between
inspections always remains low. Bridge components that have deteriorated to this extent may
no longer be performing their intended function, and remedial actions are typically planned to
address such conditions. The subjective condition rating of 3 is defined within the Recording
and Coding Guide (7) as follows:
NBI Condition Rating 3: SERIOUS CONDITION: Loss of section, deterioration, spalling or scour have
seriously affected primary structure components. Local Failures are possible. Fatigue cracks in steel or shear
cracks in concrete may be present.

12

Reliability Assessment of Bridge Elements 13

This condition description is widely understood and there is significant past experience in the
conditions warranting a rating of 3 throughout the bridge inventory. This condition description is not absolute, but provides a frame of reference for the analyst considering the likelihood
of damage occurring to a serious extent. In terms of the AASHTO Bridge Element Inspection
Guide, this condition generally aligns with elements in condition state (CS) 4, severe. (8)

2.2 Occurrence Factors


What can go wrong, and how likely is it to occur?
The first step in the reliability assessment is to address the question What can go wrong, and
how likely is it to occur? The first part of this question, what can go wrong addresses the damage
modes that affect typical bridge elements. In other words, what damage is likely to develop over
the service life of the bridge, which may result in the failure of a given element? Failure used
in this context is serious damage to the element such that its performance as intended cannot be
assured, as described in Section 2.1 (e.g., condition rating = 3 or CS = 4). For concrete elements,
spalling and cracking of the concrete is a typical damage mode. For steel elements, section loss or
cracking are typical damage modes. The second part of this question, how likely is it to occur,
describes the likelihood, or probability, of failure due to that damage mode occurring, given the
design, materials, and current condition of a bridge element. The OF categorizes this likelihood
on a qualitative scale that provides an assessment of the likelihood of serious damage, i.e., failure,
occurring.
For the assessment of bridge inspection needs, the OF is usually an assessment of the likelihood that a given damage mode will result in failure (i.e., serious condition), over a time period
of 72 months (6 years). The deterioration mechanism resulting in the damage is considered in the
assessment. In some cases the OF may be an estimate of the likelihood of a certain adverse event
occurring, such as impact from an over-height vehicle or an overload. Each damage mode or
adverse event must have a separate OF, based on the likelihood of the damage mode or the event
resulting in failure of an element during the specified time interval.
The OF describes the likelihood of failure of an element in one of four categories. The scale
ranges from remote, when the likelihood is extremely small such that it would be unreasonable
to expect failure, to high, where the likelihood of the event is increased, as shown in Table 1.
To assess the appropriate OF for a given bridge element, key characteristics, or attributes,
are considered. Attributes are characteristics of a bridge element that contribute to the elements reliability, durability, or performance. These attributes are typically well-known parameters affecting the performance of a bridge element during its service life. This includes relevant
design, loading, and condition characteristics that are known or expected to affect the durability
and reliability of the element. For example, consider the damage mode of spalling due to corrosion damage in a concrete bridge deck. A bridge deck may have good attributes, such as being
in very good condition, having adequate concrete cover, epoxy-coated steel reinforcing, and
minimal application of de-icing chemicals. Given these attributes of the deck, it may be very
Table 1. OF rating scale for RBI.
Level
1

Category
Remote

2
3
4

Low
Moderate
High

Description
Remote likelihood of occurrence,
unreasonable to expect failure to occur
Low likelihood of occurrence
Moderate likelihood of occurrence
High likelihood of occurrence

14 Proposed Guideline for Reliability-Based Bridge Inspection Practices

unlikely that severe damage (i.e., failure) would occur in the next 72 months. This is based on
the rationale that the deck is presently in good condition, and has attributes that are well-known
to provide resistance to corrosion damage. As such, an OF of Low or Remote might be used
to describe the likelihood of failure due to this damage mode. Alternatively, suppose the deck is
in an environment where de-icing chemicals are frequently used, the reinforcement is uncoated,
and the current rating for the deck is a 5, Fair Condition, indicating that there are signs of distress in the deck. Based on this rationale, the likelihood of serious damage developing would be
much greater, resulting in an OF rating of Moderate or High. Past experience with decks of
a similar design, characteristics of the specific operating environment, and attributes of the deck
are combined with engineering judgment and used to support the assessment of the specific OF
for a given deck. Methodologies for determining credible damage modes and their associated
attributes are included in Appendix A.
Certain key attributes will ideally be identified as part of criteria for reassessment of bridge
inspection requirements, following subsequent RBIs. These key attributes are typically associated with condition, which may change over the service life of the bridge as deterioration occurs.
When changes in these condition attributes cause a change in the likelihood of a given damage
mode resulting in failure (i.e., the OF), reassessment of the inspection requirements is necessary.
Deterioration rate data, trends, and theoretical models can be used to support the categorization of the OFs by providing insight regarding the average, typical, or expected behavior of
elements of a similar design. Transition probabilities, Weibull statistics, or regression trends,
developed based on past inspection results, can provide insight into the anticipated behavior
of a group of similar bridge elements. Care should be taken to ensure that the bridge elements
being assessed have similar or the same attributes as those represented by the data. Theoretical
models may also be used to support these assessments. However, the complexity and variations
in the operational environment, construction variability, and current condition can be difficult
to capture in these models. Results need to be verified using engineering judgment.

2.3 Assessment of Consequences


What are the consequences?
The second factor to be assessed within an RBI process is the Consequence Factor, CF, a
categorization of the likely outcome presuming a given damage mode were to result in failure
of the element being considered. The assessment of consequence is geared toward assessing
and differentiating elements in terms of the consequences of the assumed element failure.
It should be noted that failure of an element is not an anticipated event when using an RBI
approach, rather the process of assessing the consequences of a failure is merely a tool to rank
the importance of a given element relative to other elements for the purpose of prioritizing
inspection needs.
The CF is used to categorize the consequences of the failure of an element into one of four categories, based on the anticipated or the expected outcome. Failure scenarios are considered based
on the physical environment of the bridge, typical or expected traffic patterns and loading, the
structural characteristics of the bridge, and the materials involved. These scenarios are assessed
either qualitatively, through necessary analysis and testing, or based on past experience with similar
failure scenarios. The four-level scale used to assign the CF is shown in Table 2. The CF ranges from
low, used to describe failure scenarios that are benign and very unlikely to have a significant effect
on safety and serviceability, through catastrophic scenarios, where the threat to safety and life is
significant. Thus, both short-term (generally safety related) and long-term (generally serviceability
related) consequences can be considered.

Reliability Assessment of Bridge Elements 15


Table 2. CFs for RBI.
Level

Category

Consequence
on Safety

Consequence on
Serviceability

Low

None

Minor

Moderate

Minor

Moderate

High

Moderate

Major

Severe

Major

Major

Summary Description
Minor effect on serviceability,
no effect on safety
Moderate effect on serviceability,
minor effect on safety
Major effect on serviceability,
moderate effect on safety
Structural collapse/loss of life

In assessing the consequences of a given damage mode for a given element, the RAP must establish which outcome characterized by the CFs in Table 2 is the most likely. In other words, which
scenario does it have the most confidence will result if the damage were to occur. Using the illustration of brittle fracture in a girder, it is obvious that the most likely consequence scenario would
(and should) be different for a 150-foot span two-girder bridge than for a 50-foot span multigirder bridge. For the short-span, multi-girder bridge, an engineer may state with confidence
that the most likely consequence scenario is high or moderate and that the likelihood of
severe consequences is very remote for a multi-girder bridge, based on his or her experience
and the observed behavior of multi-girder bridges. For the two-girder bridge, the consequence
scenario is likely to be Severe. As this example illustrates, the CF simply ranks the importance of
the damage mode as being higher for a two-girder bridge than for a multi-girder bridge. For many
scenarios, qualitative assessments based on engineering judgment and documented experience are
sufficient to assess the appropriate CF for a given scenario; for others, analysis may be necessary
using suitable analytical models or other methods. A series of more detailed criteria for specific
elements [i.e., decks, steel girders, prestressed (P/S) girders, etc.] are provided in the Appendix B
that can be utilized during the assessment to determine the appropriate CF for a given element
failure scenario. These criteria, combined with owner-specific requirements developed in the RAP
or from other rational sources for assessing bridges and bridge redundancy, are then used to determine the appropriate CF for a given scenario.

2.4 The Reliability Assessment Panel


An important component of the analysis process is the elicitation of expert judgment regarding the likelihood of damage and the level of associated consequences. Because design features,
construction specifications and practices, materials, environment, and bridge management
strategies differ from state to state, or even within a particular state, the expert panel should be
selected keeping in mind the need to have membership which is familiar with the operational
environment of the inventory of bridges being evaluated.
The RAP typically will consist of four to six experts from the bridge-owning agency. This panel
should include an inspection team leader or program manager that is familiar with the inspection
procedures and practices, as they are implemented for the inventory of bridges being analyzed.
The team should include a structural engineer who is familiar with the common load paths and
the overall structural behavior of bridges, and a materials engineer who is familiar with the behavior of materials in the particular environment of the state and has past experience with materials
quality issues. A facilitator may be used to assist in the analysis process. The general characteristics
of members of a RAP include the following:
1. Bridge Inspection Expert: Inspection team leader or program manager that oversaw the
specific inspection process and the reports for the bridges being evaluated. This individual
should be able to represent the inspection results reported in the bridge file, understand the

16 Proposed Guideline for Reliability-Based Bridge Inspection Practices

notes and sketches included in the file, and have an understanding of the scope and the methods of the inspections used for the bridges under consideration.
2. State Program Manager or Bridge Management Engineer: Individual familiar with the
characteristics and the behavior of the bridge inventory throughout the state.
3. Bridge Maintenance Engineer: An individual familiar with the standard methods and techniques used for bridge maintenance, the level of maintenance typical for the bridges under
consideration, and the outcomes of bridge maintenance.
4. Materials Engineer: A materials engineers who is familiar with the history of materials performance within the state. This individual should be experienced with the materials historically
used within the state, be knowledgeable of any prior problems with the quality or with the
performance of the materials used, and be knowledgeable of typical deterioration patterns.
5. Structural Engineer: An engineer with sufficient training and experience to understand the
consequences, in a structural sense, of bridge element failures. For example, the structural engineer should be able to recognize the load paths in a structure and to understand the importance
of elements in the overall structural system of the bridge.
6. Independent Experts: The RAP may include independent experts, academics, or consultants
to address specific or complex damage modes, provide independent review, and/or supplement the knowledge of the panel as needed.
7. Facilitator: A RAP facilitator may be used to assist in the RAP analysis, to lead expert elicitations, and help build consensus during the analysis process.
The expert panel may also include representatives from the FHWA to monitor the process, to
fulfill oversight responsibilities, and to assist with the implementation of the methodology used
for inspection planning.

CHAPTER 3

Determination of Inspection
Interval and Scope
This section describes the process of determining the inspection interval and scope based on
the assessment completed as described in Chapter 2. This process leads to a prioritization of
inspection needs, highlights critical damage modes for bridges, and results in an RBI practice.

3.1 Inspection Interval


The inspection interval is selected based on the RAP assessment of the OFs and CFs. Once
these factors have been determined, their numerical values are used to place a given damage
mode in the appropriate location on a reliability matrix. A typical reliability matrix is shown
schematically in Figure 2. In this figure, the horizontal axis represents the CF as determined for
a particular damage mode for a given bridge element. The vertical axis represents the outcome
of the OF assessment for a given damage mode for a given element. Elements that tend toward
the upper right corner of the reliability matrix require shorter inspection intervals, and possibly
more intense inspections, than elements that fall in the lower left corner.
The matrix is utilized to determine the appropriate maximum inspection interval for a given
bridge or bridge type. These inspection intervals are determined to ensure that the probability, or
likelihood, of failure remains low during the inspection interval. The maximum inspection interval is established in order to be consistent with the assessment of the OF, as determined over the
predefined assessment interval of 72 months, as described in Section 2.2. Keeping this in mind, the
actual maximum inspection interval is determined such that the likelihood of occurrence within
the time between inspections (i.e., the inspection interval) always remains low. For example, if the
OF is low over a 72-month period, than it may be reasonable to assign the inspection interval of
72 months (ignoring the influence of consequence for the time being). However, if it were found
that the OF were high, the analysis is really indicating a failure is relatively likely to occur before
the end of the 72-month interval. Since the goal is to ensure that the possibility of failure occurring
before the end of the interval is always low, one would shorten the inspection interval, for example
to 24 months. In other words, by inspecting every 24 months, the possibility of failure occurring
before the end of the interval (now reduced to 24 months) remains low.
Obviously, the OF is not the only parameter that should be evaluated when setting the interval. The consequence of the failure must also be incorporated into the process of selecting the
appropriate interval. Using the example above, where the OF were high and the interval was
reduced to 24 months; if the consequence of that same damage was determined to be severe, it
would be appropriate to assign a shorter interval of, for example, 12 months. This provides an
extra measure of confidence and safety (i.e., a reduction from 24 months to 12 months due to
the severe nature of the consequences). Although there are many permutations of the OF and
the CF, the above illustrates the concept.
17

18 Proposed Guideline for Reliability-Based Bridge Inspection Practices

This is a relatively easy task for elements where the OF is high and the CF is severe, and hence
an interval of 12 months or less is needed. However, if the OF is remote and the CF is low, then
it would also seem reasonable and justifiable that the inspection interval should be greater
than the longest interval assumed in the OF assessment (72 months). (If the OF is remote, this
indicates the members of the RAP concluded that there is a remote likelihood of occurrence,
unreasonable to expect failure to occur in the next 72 months for this element and damage
mode.) This information, coupled with the observation that failure, should it occur, is a low
consequence, may justify the use of an inspection interval longer than 72 months.
Figure 2. Risk matrix
for determining
maximum inspection
intervals for bridges.

The actual inspection interval selected is based on the shortest inspection interval determined
from the analysis. In other words, whichever element has the shortest maximum inspection
interval, based on the likelihood of failure and associated consequence. In certain circumstances,
there may be one element of the bridge that results in a much shorter inspection interval than the
other elements of the bridge. In such a case, a different inspection interval may be identified for
that particular element, based on engineering judgment and the discretion of the bridge owner.
For most cases, multiple elements would be expected to have the same or very similar intervals,
with the shortest interval being selected for practical reasons.

3.1.1 Inspection Scope


Under an RBI practice, the inspection scope is determined from the damage modes identified through the reliability analysis. In other words, the inspection methods used are selected
based on their effectiveness and reliability for detecting the specific damage mode(s) that are
most important. Guidelines for the selection of inspection methods to be used are included in
Appendix D. In many cases, visual inspections supplemented with sounding are well-proven
approaches to detecting typical damage in highway bridges. However, in a risk-based process,
these inspections would include hands-on access to key portions of a bridge, such that damage
is effectively identified to support the RBI assessment. For example, when assessing the likelihood of fatigue cracking in a bridge, it would be necessary to know if there were currently fatigue
cracks in the bridge. Therefore, the inspection scope used to support the assessment must utilize
an approach that is capable of making that determination. This would require hands-on access
to certain locations where fatigue cracking is likely to occur. In some cases, NDE techniques are
required, often based on a limited access for visual inspection (e.g., for detecting a crack in a
bridge pin).
Based on the assessment of the OF and the CF, damage modes for a bridge are prioritized
based on the product:
IPN = OF CF
Where IPN = Inspection Priority Number. For example, if the fatigue cracking has a moderate likelihood of occurring and the consequence is severe, then the IPN would be 3 4 = 12.
If fatigue cracking were moderately likely, but the consequence were only moderate (minor
service disruption), for example, if the bridge in question is a short-span, multi-girder bridge
with known redundancy, the IPN for that damage mode would only be 3 2 = 6. This process
highlights the damage modes that are most important, that is, most likely to occur, and have the
greater associated consequences, if they did occur.
It should be noted that the calculation of the IPN for each damage mode identified in the process
does not limit the scope of the inspection to only those damage modes. However, it does provide a
simple method to prioritize damage modes that are most important, based on a rational engineering assessment that incorporates bridge type, age, design details, condition, etc., as well as the consequences of failure.

Determination of Inspection Interval and Scope 19

3.1.2Sampling
When using the RBI approach, it may be appropriate to inspect a representative sample of
a bridge element, using the inspection method identified. This can be used to reduce or limit
inspection activities that provide little or no measure of increased benefit or that introduce risks
that are unjustified. The sampling population size (number of locations or area, for example)
should reflect the nature and type of damage to be assessed through the inspection. When damage modes are expected to be widespread and relatively uniform, such as spalling in a bridge
deck, an appropriate sampling based on area may be justified. For example, inspecting 25%
of the bridge deck to assess if delaminations are present. When damage modes are isolated or
non-uniform, such as fatigue cracks, sufficient sampling must be based on analysis to identify
the location and number of inspections. Criteria and analysis supporting the sampling should
be documented.

3.1.3 Maintenance Inspections


RBIs are typically more focused and intense than calendar-based, general-condition inspections, and the maximum interval between inspections may be increased. For bridges with
extended inspection intervals, maintenance inspections may be specified periodically to ensure
the maintenance of traffic safety and to address general maintenance needs. These inspections
are typically conducted by maintenance personnel with responsibility for the maintenance of
the roadways and the bridges in the district or region where the bridge is located. The purpose
of a maintenance inspection is to:



Identify bridge maintenance needs (minor patching, clearance of debris, vegetation control, etc.).
Confirm general conditions have not significantly changed.
Monitor unreported vehicular damage to a structure.
Evaluate traffic safety issues (maintaining signage, roadway delineations, etc.).

Intervals for maintenance inspection would typically not exceed 2 years. Such maintenance
inspections may be integrated into the business practice of a district or region.

3.1.4 Initial Inspections


Initial inspections, the first inspection of a bridge following construction or reconfiguration
of a structure (e.g., widening, lengthening, supplemental bents, etc.) are required according to
AASHTOs The Manual for Bridge Evaluation (9). In addition to this initial inspection, at least
one RBI should be conducted at the interval of 24 months prior to initiating an RBI practice
utilizing an interval greater than 24 months. Newly rehabilitated bridges should also haveat
least one RBI at the interval of 24 months following rehabilitation. The purpose of these
inspections is to ensure that construction errors or deficiencies have not significantly altered
the anticipated performance, and that a thorough inspection based on the RAP analysis has
been conducted.

3.1.5 Start-Up Inspections


When initiating an RBI practice for a bridge, the first RBI should be conducted at the regular
interval for the bridge, typically 24 months under the current NBIS. This start-up RBI will implement the practice as determined through the RAP analysis. Following the start-up inspection,
the inspection results should be assessed for conformance with criteria and attributes identified
by the RAP to determine if reassessment is necessary before implementing any modifications to
the inspection interval.

20 Proposed Guideline for Reliability-Based Bridge Inspection Practices

3.1.6 Quality Control/Quality Assurance


Quality control (QC) and quality assurance (QA) processes should be employed to ensure
quality in the implementation of RBI practices. Procedures for QC could include data model
reviews, scoring and reliability factor reviews, RAP procedures, and application of inspection
intervals based on the RBI analysis. Procedures for QA could include analysis of historical bridge
performance, consistency in data models developed from the RAP analysis, and field reviews
of bridge performance under the RBI process. Additional methods for QC and QA for bridge
inspection programs are available in the literature (10).

CHAPTER 4

Establishing an RBI Program

4.1 Overview of Process


The overall process for implementing an RBI is shown schematically in Figure 3. The process
begins with the selection of a bridge or family of similar bridges to be analyzed. For the selected
bridge or bridges, the RAP identifies credible damage modes for elements of the bridge, given
the design, materials, and operational environment. Key attributes are identified and ranked to
determine the OFs, and the appropriate CFs associated with the damage modes are analyzed.
Based on the assessment of the OFs and CFs for the bridge, the inspection practice is established
including the interval and scope (procedures) for the inspection, and criteria for reassessment
of the inspection practice. The criteria for reassessment are typically based on conditions that
may change as a result of deterioration or damage, and may affect the OFs for the bridge. The
RBI practice is then implemented in the subsequent inspection of the bridge. Following the
inspection, inspection results are assessed to determine if any established criteria have been violated, or if conditions have changed that may require a reassessment of the OF. If such changes
exist, a reassessment of the OF is completed and the inspection practice modified accordingly.
If no such changes or conditions exist, the inspection practice can remain unchanged for the
subsequent inspection interval.
Using the overall process described above, bridge owners can initiate an RBI practice for bridges
in their inventory. However, the process and inspection requirements under an RBI practice may
diverge significantly from traditional, calendar-based, and uniform inspection strategies. Therefore, consideration is needed regarding the scope of initial RBI assessments, training for inspectors and RAP members, and integration with existing software and databases. The sections that
follow discuss these considerations.

4.2 Setting the Scope of the Analysis


RBI requires increased planning resources relative to calendar-based or uniform inspection
processes. An effective strategy for transitioning from a calendar-based inspection practice to
RBI is needed to facilitate the process and ensure adequate resources are available to conduct
the necessary assessments. A suitable approach for transitioning an inspection program from
a calendar-based, uniform inspection strategy to RBI is to identify those bridges where a reliability analysis can most readily be conducted and begin the process by assessing those bridges
first. These bridges may be identified by conducting a simple qualitative risk assessment of the
overall bridge inventory. This assessment should identify those bridges or family of bridges that
are of very common design characteristics, and where significant experience exists regarding the
anticipated damage and deterioration patterns. Such an assessment can be rapidly conducted
based on general bridge characteristics such as span length, bridge type, number of spans, and
21

22 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Figure 3. Flow chart showing RBI


program activities.

current condition. For those bridges where past experience is greatest, uncertainty regarding
both the development of the damage and the associated consequences is reduced. Bridges that
are more complex, suffer from advanced forms of deterioration, or have unique design attributes
require a higher level of assessment, as shown schematically in Figure 4. More data and a more
sophisticated or more specialized assessment may be required. Therefore, to initiate an RBI
practice, bridge owners can conduct a general, fully qualitative assessment of their inventory and
assign or determine the scope of the initial assessment to be conducted.
Bridges that are of common and simple design, and are in good condition, are identified for
analysis first. These bridges can be considered to be in a low risk category because they are of

Figure 4. Schematic diagram of qualitative risk


assessment for a bridge inventory.

Establishing an RBI Program 23

simple design and there is significant experience and confidence in their performance. For example, bridge owners conduct a simple analysis of their inventory to determine bridges that are
multi-girder, short span, and in generally good condition for assessment first. Reliability assessment for these bridges may be relatively simple. Conducting the reliability assessment of these
bridges first helps develop the RBI practice and develops the knowledge and experience of
the RAP members. After this analysis is completed, the assessment moves on to bridges that
are more complex, require more data for assessment, or require more sophisticated analysis to
determine the factors necessary for a reliability assessment.

4.3 Training Requirements


As noted, the inspection planning process is more involved and complex under an RBI scheme
relative to a calendar-based inspection planning process. The approach to inspection planning is
more focused on inspection needs for the individual bridge. Further, the assessment of reliability
characteristics requires an understanding of the approach and the assessment needs. Therefore,
training for both members of the RAP and for inspectors that will implement the results of the
RBI planning process will be necessary.

4.3.1 Training for RAP Members


Participants in the RAP process require training to understand the underlying philosophy
and processes involved in conducting RBI planning. This training should provide sufficient
knowledge in the theory and underlying approach to RBI planning, address methodologies for
expert elicitation, and processes for determining the OFs and CFs required for the analysis. A
full understanding of the underlying concepts and reliability theories utilized in the process is
necessary to conduct effective assessments. Facilitators that may be used to assist in the expert
elicitations and overall reliability assessment should be similarly trained.

4.3.2 Training of Inspectors


RBI assessments for inspection planning provide a prioritization of inspection needs for a
bridge based on the anticipated or expected damage modes, and the importance of that damage
in terms of safety of the bridge. Criteria developed through the RAP process identify key condition attributes used to determine the reliability of individual elements of the bridge. Inspections
are necessary that are capable of determining these conditions, and, as such, these inspections
are typically more intense than traditional inspections that are intended to report on the general
condition of bridge components. Training is therefore required in conducting an element-level
inspection to meet the needs of an RBI assessment. Assessments for detecting specific damage
modes may be more thorough than under traditional calendar-based practices. For example,
training in the detection of fatigue cracks in steel or reliable use of sounding to detect subsurface
damage in concrete may be needed. In certain cases, training for inspectors in the application
of advanced NDE technologies may be required. Training on the use of NDE technologies is
specialized in nature, and certification and training for specific NDE technologies is typically
available from commercial sources.
Training for bridge inspectors in the underlying philosophy of the RBI approach is also
needed. Appropriate implementation of the inspection prioritization developed through the
process, and an understanding of the importance of the quality of bridge inspection outcomes, is needed to implement the process and to transition from traditional inspection
approaches.

24 Proposed Guideline for Reliability-Based Bridge Inspection Practices

4.4 Software Development and Integration


The processes for assessing the OFs, such as identifying and scoring key attributes of bridge
elements, can be repetitive once established, and therefore lends itself to software implementations. Many of the attributes identified by the RAP may already be stored in existing databases
and bridge management systems. Condition attributes and screening criteria for RBI could be
implemented through existing software developed for bridge inspection and storing bridge
inspection data, or appropriate software may be developed. Therefore, the process of implementing an RBI practice can be simplified by the development of software to more rapidly
implement the methodology. Integration with existing software and databases that store relevant
information is beneficial.

References

1. American Petroleum Institute (API), API Recommended Practice 580, Risk-Based Inspection. 2002: Washington,
D.C., p. 45.
2.National Bridge Inspection Standards, in 23 CFR part 650. 2004: USA., p. 7441974439.
3.FHWA, Revisions to the National Bridge Inspection Standards (NBIS). 1988: p. 21.
4. American Bureau of Shipping (ABS), Surveys Using Risk-Based Inspection for the Offshore Industry. 2003:
Houston, TX.
5.ASME, Risk-Based Inspection: Development of Guidelines. General Document. 1992: p. 155.
6.ASME, Inspection Planning Using Risk-Based Methods. 2007: p. 92.
7.FHWA, Recording and Coding Guide for the Structural Inventory and Appraisal of the Nations Bridges,
U.S.DOT., Editor: 1995:.
8.AASHTO, AASHTO Bridge Element Inspection Manual. 2010, AASHTO Publications: Washington, D.C.
p.170.
9.AASHTO, The Manual For Bridge Evaluation. 2008, AASHTO Publications: Washington, D.C.
10. Washer, G.A., and Chang, C.A., Guideline for Implementing Quality Control and Quality Assurance For
Bridge Inspection. 2009, Transportation Research Board: Washington, D.C. p. 65.

25

APPENDIX A

Guideline for Evaluating


the Occurrence Factor
27

A 1 Introduction

28

A 2 Damage Modes

28
29

30
31
32
32
33

34
34
34
35
36
37
38

26

A 2.1 Expert Elicitation for Credible Damage Modes


A 2.2 Example of Soliciting Expert Judgment for Damage Modes

A 3 Element Attributes
A 3.1 Screening Attributes
A 3.1.1 Qualitative Assessment of Elements and Details
A 3.2 Identifying Key Attributes
A 3.3 Ranking Attributes

A 4 Occurrence Factor Assessment


A 4.1 Estimating the Occurrence Factor
A 4.2 Calibration of Scoring Regime
A 4.3 Occurrence Factor Scale Numerical Estimates
A 4.4 Use of Deterioration Rate Data
A 4.5 Use of Surrogate Data
A 4.6 Rationale and Criteria Based on Rap Assessments

Guideline for Evaluating the Occurrence Factor 27

A 1 Introduction
The Occurrence Factor (OF) is used within an RBI to estimate the likelihood of serious damage (i.e., failure) developing in a bridge element during a specified time interval, based on engineering rationale. This rationale is developed through a systematic process that considers and
documents the anticipated damage modes for bridge elements. The characteristics, or attributes,
of bridge elements that contribute to their reliability, considering the expected damage modes,
are identified. The damage modes and attributes are identified through an expert panel process
described herein, and subsequently used in a rational process that identifies those bridges with
elements that are highly reliable and durable, and those bridges with elements that are more
likely to suffer from deterioration and damage.
The overall process for estimating the OFs is as follows:
1. Identify the likely damage modes that will affect a bridge element from commonly known
damage modes, past experience, and engineering judgment.
2. Identify attributes that contribute to the reliability and the durability of the element considering the damage modes identified.
3. Rank the importance of each attributes influence on the reliability and the durability of the
bridge element.
4. Develop rationale based on the damage modes and attributes of the bridge element to estimate the likelihood of serious damage (i.e., failure) occurring during the specified interval.
An empirical scoring procedure based on the key attributes identified for a given element is
used to provide a rational method of estimating the OF. The analysis can be used to construct
criteria that can be applied to individual bridges, or groups of very similar bridges, to categorize
the likelihood of serious damage (i.e., failure) occurring in the next 72-month time frame into
one of four categories, ranging from remote to high, i.e., the OF.
The OF represents a probability of failure (POF) estimate over a time interval of 72 months.
This time period was selected based on engineering factors that included prior research, analysis
of data from the National Bridge Inventory (NBI), expert judgment, and data from corrosion
and damage models. It was also selected as a time interval for which an engineer could reasonably be expected to estimate future performance within four fairly broad categories, ranging
from remote to high, based on key attributes that describe the design, loading, and condition
of a bridge or bridge element. In addition, this time interval was selected to provide a suitable
balance between shorter intervals, when the POF could be unrealistically low due to the typically slow progression of damage in bridges, or longer intervals, where uncertainty would be
increasingly high.
The analysis provides the rationale for categorizing the OF on a rating scale from remote,
when the likelihood is extremely small such that it would be unreasonable to expect failures, to
high, where the likelihood is increased. This rating scale is shown in Table A1. In some cases,
the OF may be an estimate of the likelihood of a certain adverse event occurring that results in a
failure, such as impact from an over-height vehicle or an overload.

Table A1. Occurrence factor rating scale for RBI.


Level
1

Category

2
3
4

Low
Moderate
High

Remote

Description
Remote likelihood of occurrence,
unreasonable to expect failure to occur
Low likelihood of occurrence
Moderate likelihood of occurrence
High likelihood of occurrence

28 Proposed Guideline for Reliability-Based Bridge Inspection Practices

The following sections describe how a Reliability Assessment Panel (RAP) identifies the damage modes to be assessed, determines important attributes for each damage mode, and ranks
and scores those attributes to support assessment of an individual bridge or families of bridges
of nearly identical attributes, damage modes, and design. The RAP is an expert panel assembled
by the bridge owner as described in section 2.4 of the main report.

A 2 Damage Modes
The first step in the process is to answer the question What can go wrong? For most common
bridges, the damage modes that affect the bridge are well known. Spalling and cracking of the
concrete as a result of corrosion, or section loss and fatigue cracking in steel elements, are typical examples. The RAP, through a consensus process, develops a listing of the credible damage
modes for the elements of a bridge or a family of bridges being assessed. A credible damage
mode is one that could reasonably or typically be expected to occur during the service life of the
bridge element. Current and past research and experience should be considered in developing
the listing. An expert elicitation process described in section A 2.1 may be used to identify the
typical damage modes for consideration. This process may also be used to identify unusual or
uncommon damage modes that may be relevant for a particular bridge inventory. Table A2 lists
damage modes that may be identified by the RAP, as examples to illustrate typical damage modes
for several common bridge elements.

A 2.1 Expert Elicitation for Credible Damage Modes


In many cases, the credible damage modes for a given bridge element may be readily identified from past experience and engineering knowledge. In other cases, it may be necessary for the
RAP to form a consensus on the credible damage modes for a given element. To identify damage
modes that are specific to the type of bridge and elements being considered, the RAP can utilize
a process to elicit the expert judgment of the panel based on their experience and knowledge.
The process is an expert elicitation of judgments from the panel that consists of the following:
1. Identify the element scenario: The first step in the process is to frame the problem for the
panel. This includes describing the element under consideration, including the material
and known design parameters. The operational environment for the element should also be
described, such as the environment and loading, especially if the operational environment
is atypical or unique. For example, if the element under consideration is a concrete beam
located in an aggressive coastal environment.

Table A2. Typical damage modes for common


bridge elements.
Element
Steel Girder

Prestressed Girder

Piers and Abutments

Damage Modes
Corrosion damage/section loss
Fatigue cracking
Fracture
Impact damage
Corrosion damage (spalling/cracking)
Strand fracture
Shear cracking
Flexural cracking
Impact damage
Corrosion damage (spalling/cracking)
Damage to bearing areas
Unexpected settlement/rotation

Guideline for Evaluating the Occurrence Factor 29

2. Identify damage modes: The facilitator poses a question to the RAP such as: The inspection
report indicates that the element is rated in serious condition. In your expert judgment, what
is the most likely cause (i.e., damage mode) that has produced/resulted in this condition?
This question is intended to elicit from the panel a listing of damage modes that are likely to
occur for the element.
Each expert is asked to independently list the damage modes he/she judges are most likely to
have caused the element to be rated in serious condition. The expert records each damage mode
he/she identifies, and provides an estimate of the relative likelihood that the damage mode was
the cause. This is done by assigning relative probabilities to each damage mode, typically with
a minimum precision of 10% (the sum of the ratings should be 100%). The expert notes any
supporting rationale for their estimate. The individual results from each member of the RAP are
then aggregated to evaluate consensus among the panel on the most likely damage modes for
the element. An iterative process may be necessary to develop consensus on the credible damage modes for a given bridge element. However, for most elements, the damage modes are well
known and consensus can be reached quickly.

A 2.2 Example of Soliciting Expert Judgment for Damage Modes


This section provides an example of the process for eliciting expert judgment from the RAP
for a typical bridge element. In this example, the RAP is provided with the following description for a steel bridge member: The element under consideration is a painted, rolled steel girder
in a simply supported, multi-girder bridge with a typical span length, in a moderate environment.
If you were told this girder is rated in serious condition, what would be the most likely cause of this
condition?
Each member of the RAP is then asked to list the damage modes that they identify as the most
likely causes (e.g., cracking, section loss) for the member condition, and estimate its relative
likelihood of being the cause, relative to other damage modes they identify. The results of this
independent exercise are then aggregated as shown in Table A3, showing illustrative results from
a six member RAP team assessing the given element scenario.
Following the independent elicitation, the panel discusses the results of the assessments. Any
damage mode with an average score of less than 10% may be assessed to determine if that damage mode is credible for the given scenario. Rationale for inclusion or exclusion of the particular
damage mode should be recorded. Any damage modes with variance of >20% from the average
are also discussed, and RAP members are provided an opportunity to revise their individual
ratings based on the discussions.
In this steel girder example, the panel considers the damage mode of corrosion damage/section
loss to be most likely to have resulted in severe damage to the steel girder. Less likely damage
modes include fatigue, overload, and impact damage. Each credible damage mode identified will
be assessed by the RAP to determine its OF.

Table A3. Expert elicitation of damage modes for steel girders.


Damage
Corrosion/
section loss
Fatigue
Overload
Impact
Sum

Expert 1

Expert 2 Expert 3

Expert 4

Expert 5

Expert 6

Average

60%

60%

50%

50%

70%

50%

57%

30%
10%
0%
100%

30%
10%
0%
100%

30%
10%
10%
100%

20%
20%
10%
100%

10%
20%
0%
100%

20%
20%
10%
100%

23%
15%
5%
100%

30 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table A4. Example damage modes for a steel girder bridge.
Element
Bare Concrete Bridge Deck

Steel Girder

Piers and Abutments

Damage Modes
Spalling resulting from steel corrosion
Widespread cracking
Rubblization of concrete due to freeze/thaw
damage or ASR
Corrosion damage
Fatigue cracking
Fracture
Impact damage
Spalling resulting from corrosion
Damage to bearing areas
Unexpected settlement/rotation

The elicitation process is repeated for each key element of the bridge to develop a listing of
damage modes to be considered in the analysis. For example, considering a typical steel girder
bridge with a bare concrete deck and concrete piers and abutments, damage modes for each
element of the bridge that might be identified by a RAP are shown in Table A4. For the deck
in this illustration, the most common damage mode is identified as spalling of the deck due to
corrosion damage of the reinforcing steel; widespread cracking, and damage due to alkali-silica
reactivity (ASR) and/or freeze-thaw cycles. For the steel girder, corrosion damage (section loss)
is identified as the most likely damage mode; fatigue cracking, fracture, and impact are also
identified by the RAP. For the piers and abutments, damage modes included corrosion damage
that results in spalling, damage to the bearing areas (beams seats, for example), and unexpected
settlement or rotation. Such a listing is developed through a consensus process by the RAP for a
specific bridge and element types under consideration, as previously discussed.
Once this listing of damage modes has been identified, the next step in the process is to identify key attributes that contribute to the reliability and durability of the element, considering
these damage modes.

A 3 Element Attributes
Attributes are characteristics of a bridge element that affect is reliability. These attributes
are typically well-known parameters affecting the performance of bridge elements during their
service lives. For example, bridge elements can have good attributes that are known to provide
good service-life performance. A bridge deck can have good qualities such as having adequate
concrete cover and use of epoxy-coated reinforcing steel for corrosion resistance. Alternatively,
bridges may have qualities or attributes that contribute to more rapid deterioration or increased
likelihood of damage. Using the concrete deck example, heavy use of de-icing chemicals, minimal concrete cover, and unprotected reinforcement would be examples of attributes that contribute to more rapid deterioration. For a steel girder, fatigue-prone details may be an attribute
indicating increased likelihood of damage. The identification of key attributes is simply a listing
of these attributes and a relative ranking of their importance in terms of the reliability and the
durability of the element.
These attributes can be generally grouped into three categories: Design, Loading, and Condition attributes. Design attributes are characteristics of a bridge element that are part of the elements design. Design attributes are usually intrinsic characteristics of the element that do not
change over time, such as the amount of concrete cover or material of construction [concrete,
high performance concrete (HPC), etc.]. In some cases, preservation or maintenance activities

Guideline for Evaluating the Occurrence Factor 31

that contribute to the durability of the bridge element may be a design attribute, such as the use
of penetrating sealers as a preservation strategy.
Loading attributes are characteristics that describe the loads applied to the bridge element that
affect its reliability. This may include structural loading, traffic loading, or environmental loading. Environmental loading may be described in macro terms, such as the general environment
in which the bridge is located, or on a local basis, such as the rate of de-icing chemical application
on a bridge deck. Loading attributes describe key loading characteristics that contribute to the
damage modes and deterioration processes under consideration.
Condition attributes are characteristics that relate to the current condition of a bridge or a
bridge element. These can include the current element or component level rating, or a specific
condition that will affect the reliability of the element. For example, if the damage mode under
consideration is concrete damage at the bearing, the condition of the bridge joint may be a key
attribute in determining the likelihood that severe corrosion will occur in the bearing area.
Relevant attributes are identified for the damage modes and underlying deterioration mechanisms determined by the RAP. In many cases, attributes are well-known characteristics of bridges
and bridge elements that contribute to the reliability and durability of the elements. However,
because bridge designs, environments, and management policies differ, attributes and their
relative importance may also differ between bridge owners. Therefore, it is necessary that the
RAP identify those attributes that contribute most significantly, including any special or unique
attributes that might contribute significantly (either positively or negatively) to the likelihood of
damage for bridges in their inventory. Attributes that are not relevant or do not have significant
impact on durability and reliability should not be included in the analysis.

A 3.1 Screening Attributes


Screening attributes can be used to quickly identify bridges or elements that should not be
included in a particular analysis, either because they already have significant damage or they
have attributes that are outside the scope of the analysis being developed. Screening attributes
are typically attributes that:
Make the likelihood of serious damage occurring very high.
Make the likelihood of serious damage occurring unusually uncertain.
Identify a bridge with different anticipated deterioration patterns than other bridges in a group.

Once the attribute listing has been completed, attributes that match these criteria can be
identified. The RAP should identify the appropriate value or condition for the attribute to
use as a screening tool. In any scoring scheme there is the possibility, and hence a concern,
that the value of key attributes can be diminished when the scoring for all of the relevant attributes are combined. Screening attributes are useful to ensure key conditions are identified, to
address this concern.
For example, if considering the likelihood that the steel bridge will suffer corrosion damage that reduces its rating to a 3, and the current rating is 4, the RAP may consider that such
condition indicates that there is a high likelihood of further damage developing over the next
72-month period, regardless of other attributes. In such a case, the analysis can move forward to
an assessment of the consequences without assessing the specific attributes of the element, since
the likelihood has already been assessed to be high.
Design features may be useful as screening criteria, particularly if the features result in the
likelihood of serious damage being unusually uncertain. For example, for bridges that possess
details susceptible to Constraint-Induced Fracture (CIF), there is a high potential for sudden

32 Proposed Guideline for Reliability-Based Bridge Inspection Practices

brittle fracture. For fracture-critical bridges in particular, inspection will provide no protection
as the CIF occurs without any warning and before any detectable cracks are observed. Hence,
it would be prudent to screen these bridges from the analysis, because the likelihood of serious
damage is unusually uncertain. Another strategy, such as retrofitting the critical details, should
be performed to ensure safety.
Another example would be to screen steel beam elements in bridges that have open decking.
Since the open decking allows drainage directly onto the steel beams, the deterioration of these
bridges would not be similar to steel beams with typical concrete decks. Therefore, it would be
prudent for these bridges to be screened from the analysis of steel beam bridges, as they may
require separate analysis. It may be appropriate to treat these bridges as a separate group, developing the analysis to consider key attributes of those bridges with open decking.
In some cases, it may be more practical to screen bridges from the analysis entirely through a
qualitative reliability assessment of the overall inventory, as described in the following section.
A 3.1.1 Qualitative Assessment of Elements and Details
A simple qualitative assessment can also be used early in the RAP process to identify appropriate families or groups of bridges to be analyzed. This tool can be used to separate potentially
problematic details or elements that may require more in-depth analysis. These elements may
include, for example, rocker bearings in long-span bridges, modular expansion joints, or other
details that have the potential to affect the reliability of a bridge uniquely. The qualitative assessment uses a simple three-level scale, as shown in Table A5. This tool can be used to perform an
assessment of a bridge inventory and sort bridges that include attributes that are perceived to
have low reliability or require special analysis. The assessment is useful for identifying bridges
that can be easily assessed from those for which more detailed or individual assessments may
be required. For example, assume the RAP is going to assess multi-girder rolled beams, but it
considers those beams with rocker bearing to require special analysis and to potentially have low
reliability (relative to bridges with other bearing types); these bridges are simply screened from
the process using the qualitative assessment, such that the balance of the bridges in that family
can be assessed appropriately. A separate analysis that addresses this specific attribute can then
be developed, if necessary.
This qualitative screening process would typically be used early in the reliability assessment
process to identify an appropriate family or group of bridges and make assessments more efficient.

A 3.2 Identifying Key Attributes


Attributes can be identified generally through a variety of means such as past performance,
experience with the given bridge element, previous and contemporary research, analysis of historical performance, etc. While there are potentially many attributes that contribute, in some
way, to the durability and reliability of a bridge element, it is necessary to identify those attributes
that have the greatest influence on the future performance of an element. Key attributes for a

Table A5.Qualitative
reliability scale for
screening details.
Relative Reliability
High
Moderate
Low

Guideline for Evaluating the Occurrence Factor 33


Table A6. Attributes related to the damage mode
of corrosion for a steel girder.
Design Attributes
Deck
Joints/Drainage
Built-Up Members

Loading
Attributes
MacroEnvironment
MicroEnvironment

Condition Attributes
Existing Condition
Joint Condition

Deck Type

Maintenance Cycle

Age/Yr of
Construction

Condition History/
Trend
Debris Accum.

given damage mode can be identified through expert elicitation of the RAP. For example, the
facilitator could ask the following question pertaining to a particular damage mode, X:
Consider damage mode X for the subject bridge element. If you were asked to assess the likeli-

hood of serious damage occurring in the next 72 months, what information would you need
to know to make that judgment?
The resulting input from the RAP can be categorized appropriately and ranked according to
the relative importance of the attribute for predicting future damage for the identified damage
mode and element. Rationale for each attribute should be documented. Many of the most common attributes are described in Appendix E, and can be documented by reference. For attributes
not included in Appendix E, a brief summary of the rationale for the attribute should be developed and recorded by the RAP.
As an example, Table A6 illustrates typical attributes identified by a RAP for corrosion damage
on a steel girder element. Based on an expert elicitation, the primary attributes that contribute to
the likelihood of serious corrosion damage developing for a steel girder bridge element include
design attributes, loading attributes, and condition attributes, as shown in the table. The rationale for these attributes is relatively simple and straightforward. For example, the presence of
deck joints and the quality of the drainage system may indicate whether or not the bridge has
deck drainage that is likely to spill de-icing chemicals directly onto the steel girder, thereby resulting in an increased likelihood of corrosion occurring. Built-up members are more likely to suffer
crevice corrosion and would therefore be more likely to suffer serious corrosion damage than a
rolled or welded section. The attribute of deck type considers if there is open decking that allows
de-icing chemicals to drain directly onto the girder, thereby increasing the likelihood of corrosion damage, etc. These attributes are identified by the RAP by applying common engineering
knowledge to develop criteria from which a steel bridge element can be assessed to determine if
it is likely to suffer serious corrosion damage, or if corrosion damage is unlikely. Elements that
have little exposure to de-icing chemicals, are in mild environments, and are currently in good
condition may be unlikely to develop serious corrosion in the near future. Conversely, a steel
element with active corrosion present, which is in an aggressive environment, and/or is exposed
frequently to de-icing chemicals, is more likely to develop serious corrosion damage.

A 3.3 Ranking Attributes


Once the key attributes have been identified, the attributes are ranked on a simple three-level
scale according to their importance in assessing the reliability of a bridge element. The ranking
is based on the consensus of the RAP. This scale, shown in Table A7, is used to rank a particular
attributes importance as high, moderate, or low. Once ranked, the attributes are assigned a point
value corresponding to their importance, to be used in the attribute scoring methodology that

34 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table A7. Ranking scales for
key attributes.
Ranking Descriptor

Total Points

High

20

Moderate

15

Low

10

supports the RAP assessment of the OF. For attributes that are ranked with high importance, a
scale of 20 points can be assigned, 15 points for an attribute that has a moderate importance, and
10 points for an attribute that plays a minor role, but is still an important indicator. For example,
for the corrosion of a steel beam, a leaking joint which results in drainage of de-icing chemicals
directly onto the superstructure is highly important in assessing the likelihood of serious corrosion damage occurring. Therefore, this attribute would be assigned a 20 point scale. Age of
the structure contributes to the likelihood of corrosion damage, but to a much lesser extent,
relatively, such that it would have 10 points allocated. Maintenance cycle, built-up members,
and debris build-up are moderate indicators; these may be assigned a point scale of 15 points.
Once the importance of the attribute is identified, different conditions or situations may be
described to distribute points appropriately based on the engineering judgment of the RAP.
Again, a simple high-, moderate- and low-ranking model should be used to distribute scores
among different conditions or situations that are appropriate for a given attribute. Depending
on the number of different conditions or situations, scoring may be distributed over two, three,
or four different levels for a given attribute. Using a joint as an illustration, if the joint is leaking
or can reasonably be expected to be leaking, it will have the highest effect and might be scored the
full 20 points. If the joint is debris-filled or exhibiting moderate leakage, a score of, for example,
15 points may be appropriate; if there is a joint, but it is not leaking, a score of 5 points may be
assigned. If the subject bridge is jointless, a score of 0 points may be used. The distribution of
scoring for a particular attribute is determined by the RAP. Numerous examples for scoring various attributes are included in the Attribute Index and Commentary located in Appendix E. The
RAP should assess if the suggested scores in Appendix E are appropriate, based on the characteristics of the bridges being assessed, and assign appropriate scoring regimes for attributes selected.

A 4 Occurrence Factor Assessment


A 4.1 Estimating the Occurrence Factor
Once the appropriate attributes have been identified and ranked for a given element, the
attributes are used to estimate the appropriate OF for a that element. A simple scoring procedure
is developed to evaluate the reliability characteristics of a given element based on the attributes
and their relative ranking, as described above. The developed scoring procedure provides a data
model that is used to assess the OF. Attributes scoring sheets may be used to record the relative
scoring of the attributes for a given element, or the data model may be implemented through
suitable software. Illustrative examples are included in Appendix F.

A 4.2 Calibration of Scoring Regime


Once the appropriate attributes are selected and ranked, the overall outcome of the scoring
procedure (i.e., data model) should be tested to ensure results are adequate for categorizing the
subject elements. In some cases, the weighting of particular attributes may need to be increased

Guideline for Evaluating the Occurrence Factor 35

or decreased to provide suitable results. Since operational environments and design and construction practices vary, rankings for attributes and associated values may need to be adjusted.
When a large number of attributes are identified, the relative weight of the most important
attributes becomes diminished relative to the overall scoring, and may need to be adjusted to
appropriately characterize the anticipated reliability of the element. Screening attributes can also
be used for this purpose. Sensitivity studies and Monte Carlo simulation may also be used to
assess the relative weights designated for attributes and calibrate the scoring regime developed.
The effectiveness and accuracy of the scoring regime developed can also be evaluated using
back-casting, a process for analyzing historical inspection records to verify the effectiveness of
the data model (i.e., attributes and scoring) developed. In a back-casting assessment, the attributes and scoring regime are applied to historical inspection records to assess their effectiveness
for identifying the likelihood of serious damage occurring.
Regardless of the method(s) used to calibrate the data model, engineering judgment should
be used to verify the adequacy of the data model developed.

A 4.3 Occurrence Factor Scale Numerical Estimates


The OF is a qualitative assessment of the likelihood of failure occurring during the next
72 months. Four categories are used to characterize the likelihood considering a particular element and damage mode. Table A8 includes numerical ranges that could be used to describe the
OF scale quantitatively. Such numerical values provide ranges or target values for the qualitative
rankings that could be used to map quantitative data, if these were available, to the qualitative
rating scales. Failure of a bridge element is a relatively rare event, and design and construction details vary widely. As a result, relevant and verifiable frequency-based probability data are
scarce. In some cases, modeling may be used to provide an estimate of a particular failure frequency or probability. Probabilistic models or assessments may also be developed for a particular bridge element or elements. The numerical values shown in Table A8 are target values that
could be used to map these data or models to the qualitative scales used for analysis. Providing
a quantitative estimate of the OFs allows for the data from the probabilistic analysis to be incorporated directly in the reliability-based bridge inspection practice. These numerical categories
can also provide a framework for future development of models or data derived from analysis of
the deterioration patterns in a particular bridge inventory.
An estimate of a particular damage mode having a low likelihood is somewhere between
1/1,000 and 1/10,000. Although the quantitative probability is not necessarily known, engineering judgment supported with an evaluation of the reliability characteristics of the elements is adequate to differentiate between different categories: for example, a likelihood in
the low category, where the chances are 1/1,000 or less, versus something in the moderate

Table A8. Occurrence Factor categories


and associated interval estimates.
Level

Category

Description

Likelihood

Remote

Remote probability of
occurrence, unreasonable to
expect failure to occur

1/10,000

Low

Low likelihood of occurrence

Moderate

Moderate likelihood of
occurrence

1/1,0001/10,000
1/1001/1,000

High

High likelihood of occurrence

>1/100

36 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table A9. Percentage estimates for Occurrence
Factor ratings.
Qualitative Description
Remote

Expressed as a Percentage
0.01% or less

Low

0.1% or less

Moderate

1% or less

High

>1%

category, where the likelihood in less than 1/100 but greater than 1/1,000. Estimates from
deterioration rate information or from statistical modeling can also be used to support the
categorization of the OF.
The quantitative description can be also be used as a vehicle for expert elicitation by using
common language equivalents for engineering estimates. For example, if you asked an expert
to estimate the probability of serious corrosion damage (widespread spalling, for example) for
a particular bridge deck given its current condition, a common engineering response might
include a percentage estimate, for example, less than 0.1% chance or less than 1 in a thousand.
This estimate can then be mapped to the qualitative scale as being low. Such estimates are
typically very conservative, particularly for lower, less likely events. For engineering estimates
of the likelihood of a failure occurring for a given bridge element, the qualitative scale can be
interpreted as shown in Table A9.

A 4.4 Use of Deterioration Rate Data


Data on the previous performance of bridge elements can provide some insight into the likelihood of damage occurring for a bridge element. Such data can provide supporting information
for decision making regarding the appropriate OFs for a family of similar bridge element types.
The user is cautioned that deterioration rate data records only historical events that may not
reflect the rate or likelihood of future events. For example, a state may have never had corrosion damage occur in prestressing tendons; however, this provides little insight into how likely
it is that tendon damage will occur in the future. It may be that the population of bridges from
which the data is obtained has simply not reached the age where tendon damage would become
apparent. Further, deterioration rate data based on condition states or condition ratings may
provide little insight into the deterioration mechanisms that caused the condition states or
ratings to change. Caution and careful judgment should be used in determining the relevance of
the deterioration data to the particular bridge under consideration. Considerations for utilization
of deterioration rate data include:
Similarity of Operational Environment: The RAP should consider if the particular bridge

under consideration shares the same operational environment as the elements from which
data were obtained. Key elements of the operational environment include the average daily
traffic (ADT), average daily truck taffic (ADTT), macro-environment of the bridges (severe
environment vs. benign environment), micro-environment (salt application, joint and drainage conditions, exposure to overspray), and typical maintenance and management.
Similarity of Key Attributes: Key attributes that affect the damage modes and mechanisms
for the bridge element should be similar for the bridge under consideration to those from
which deterioration rate data were obtained. This may include materials of construction,
design attributes, and condition attributes. Quality of construction and years in service may
also be a factor.

Guideline for Evaluating the Occurrence Factor 37

Deterioration rate data typically describe the mean or average behavior of the bridge element
based on the observed behavior of a population of similar elements. Statistical descriptors of the
dispersion of the data, such as the standard deviation, may be provided and then used as indicators of the variability of the data. Applying such data to a specific bridge assumes that the specific
bridge has the same design, operational environment, and attributes as those in the larger population from which the statistics were derived. Attributes identified through the RAP process may
be used to judge if a particular bridge or family of bridges could be expected to perform above
the average or mean, or below.
Statistical data from a bridge management system or other databases can also be used to inform
this process if it is available. This data can be useful in determining the damage modes and the
overall deterioration behavior of similar bridge elements in the past. However, this data should not
be used exclusively because past experience does not necessarily indicate what would occur in the
future. Therefore, it is important that the RAP utilize their collective engineering judgment, experience, and rationale for identifying and assessing damage modes that can affect bridge elements.
Lastly, when using such data, one would have to decide which data to use: the mean, or say,
two standard deviations below the mean. If the mean is used, there may be a 50-50 chance that the
bridge being assessed will deteriorate more quickly than predicted by using the mean deterioration data. However, using some confidence limit, say 2 standard deviations from the mean, may
be overly conservative and result in all bridges, good or bad, having unrealistically high estimates
of likelihood. Thus, using such data without the ability to also consider or incorporate specific
information (condition, design data, details, etc.) from the bridge under consideration must be
done with caution, and with a full understanding of the ramifications of such an approach.

A 4.5 Use of Surrogate Data


For many bridges, the use of surrogates for the attributes identified in the reliability analysis
may be considered to improve the efficiency of the analysis for larger families of bridges. As used
herein, surrogate refers to specific data that can be used to either infer or determine another
piece of information that is required for the reliability assessment. For example, assume a fracture critical bridge was designed and built in the year 2000, which is well after the implementation of the AASHTO/AWS Fracture Control Plan. This information can be used to determine
that the steel must at least meet certain minimum toughness requirements, and the bridge meets
modern fatigue design requirements. Note that this was determined only from the date of construction and with no detailed review of the design calculations or specifications.
As stated, the use of surrogates is particularly attractive when identifying and assessing a family of bridges. Design and loading attributes identified by the RAP are typically static in nature,
that is, they do not change over time. The condition attributes will typically change over time,
as damage accumulates and deterioration mechanisms manifest. However, when elements are
in generally good condition, specific condition attributes identified by the RAP may not require
individual assessment for each bridge or family of bridges; the previous inspection results can
simply be used as a surrogate for the individual attributes. This will typically allow for larger
groups of bridges of similar design to be grouped into a particular inspection interval, based on
the criteria developed by the RAP. For example, again considering steel bridges built to modern
design standards, it is known that the design attributes that would increase the likelihood of
fatigue cracking and fracture have been mitigated through improvements in the design, fabrication, and construction process. The condition attributes that are required to assess the reliability
of the element would include the presence of fatigue cracks due to out-of-plane distortions,
fatigue cracking due to primary stresses, and corrosion damage. However, if the component
rating is 7, in good condition according the NBIS scale, or CS 1 in an element-level scheme, the
existing ratings can be used as a surrogate for the condition attributes. Note: This assumes the

38 Proposed Guideline for Reliability-Based Bridge Inspection Practices

inspection result is from an RBI procedure, i.e., the inspection was capable of identifying the necessary condition attributes. This allows all bridges that are of this same rating (and similar design,
loading, and condition attributes) to be treated collectively in a process that is data-driven and
does not require much detailed analysis of individual bridges. If the condition rating or condition state changes, then the bridges can be reevaluated, according to the RAP criteria. If the
condition does not change between periodic inspections, reassessment may not be necessary.
It is important to note that this process is significantly different than assigning an inspection
interval based simply on the current condition of the bridge, for example, deciding to inspect all
steel bridges with a rating of 7 on a longer interval than all of those rated a 6. The RAP analysis
forms a rationale that identifies not only the current condition attributes that affect the reliability
of the element, but also the design and loading attributes of the bridge or bridge element that
affect the potential for damage to occur. This RAP evaluation forms an engineering rationale for
the decision-making process that considers not only the condition of the element, but also the
damage modes and the potential for that damage to occur.
For element-level inspection schemes, the attributes identified by the RAP may map directly
to an element and element condition state. For example, consider that the RAP identifies leaking joints as an attribute driving the likelihood of section loss in the bearing area of a steel beam.
The element condition state (joint leaking) is recorded in the inspection process and can be used
as a criterion for that attribute score. In some cases, all of the attributes identified by the RAP
as being critical to the likelihood of failure of an element may be included in a comprehensive
element-level inspection process, in other cases, they may not.
For NBI-based inspection schemes, attributes identified by the RAP may map to sub-element
data collected in addition to the required condition ratings for the primary components of the
bridge. These data could be used if it is collected under a standardized scheme for rating and
data collection for the sub-elements. For the primary components, the generalized nature of the
component rating makes this more difficult for specific attributes.
Mapping of the attributes from the RAP analysis to the elements, sub-elements, or element
condition states should not be performed until the RAP analysis has been completed independently. In some cases, the RAP analysis may identify attributes or factors not presently included
in the available data, and these data may need to be obtained from other sources. For example, for
the case of fatigue cracking in a steel beam, element condition states would indicate fatigue cracking, but not the presence of fatigue sensitive details, i.e., the potential for cracking may be high,
even though no cracking is currently present. This is an important consideration in the assessment
of appropriate scope and interval of inspection. This data may be readily available in the bridge file,
or may need to be ascertained from design plans, records, or other data on the bridge design. In
any case, the RAP analysis shall not be constrained by the data presently available; the RAP should
identify what data is needed and then assess if that data is readily available. In some cases, additional data may need to be collected to support the analysis.

A 4.6 Rationale and Criteria Based on RAP Assessments


The RAP assessment for a given bridge or a family of bridges provides an engineering rationale for decision making regarding the appropriate inspection interval and scope. The effects of
design, loading, and condition attributes on the potential for failures are considered and documented through the process. For most bridges, the design attributes and loading attributes will
not change over time. The RAP assessment should include criteria for modifying the selected
inspection interval, and/or for reassessment of a bridge, based on the results of the RBI. These
criteria will typically be based on the condition attributes identified during the RAP assessment.
If loading conditions change significantly, reassessment may be necessary.

APPENDIX B

Guideline for Evaluating the


Consequence Factor
40
41

41
42
42
43
43
43
43
46
46
46
48
48
48

B 1Introduction
B 1.1 Definitions

B 2 General Descriptions of Consequence Scenarios


B 2.1 Low Consequence Event
General Description
Requirements for Selection
B 2.2 Moderate Consequence Event
General Description
Requirements for Selection
B 2.3 High Consequence Event
General Description
Requirements for Selection
B 2.4 Severe Consequence Event
General Description
Requirements for Selection

50

B 3Use of Expert Elicitation for Determining


the Consequence Scenario

50

B 4References

39

40 Proposed Guideline for Reliability-Based Bridge Inspection Practices

B 1 Introduction
Within an RBI, the Consequence Factor (CF) is used to categorize the outcome or the result
of the failure of a bridge element due to a given damage mode. For example, brittle fracture is
one of the key damage modes pertaining to steel bridges. Should brittle fracture of a girder occur,
the next logical question becomes, what is the consequence? This would obviously depend on
the specific scenario for the fracture. If the member were classified as fracture critical, such an
event may be catastrophic, or one that would be considered to be a severe consequence. However, if the girder were one member of a multi-girder short-span bridge, the consequence of that
fracture would likely to be much less serious, perhaps requiring a lane closure or even temporary closure of the bridge, or a high consequence. (Multi-girder bridges described herein are
bridges with four or more main load bearing members.) In fact, in some cases, such an event
may only have moderate consequences.
The CF is used to categorize the consequence of failure of a bridge element into one of four
broad categories: Low, Moderate, High, and Severe. Table B1 indicates the general descriptions
for each of the CF categories used for the RBI assessment. The general descriptions are indicated in terms of safety and serviceability of the bridge, graduated with qualitative descriptions.
Both long- and short-term consequences should/may be considered.
To assess the appropriate category for a particular element and damage mode, typical scenarios
or outcomes of a failure must be considered. In some cases, there may be a single scenario that
could result from the failure of an element; in other cases, more than one possible scenario needs
to be considered. Using the example of brittle fracture of a single beam in a multi-girder, shortspan bridge as noted above, it is unlikely that the result from a brittle fracture is a low consequence,
which has a minor effect on serviceability and no effect on public safety. It is much more likely that
such a fracture may have a moderate consequence, which has a moderate effect on serviceability
and a minor effect on public safety. It is also possible that the fracture will have a high consequence,
which has a major effect on serviceability and a moderate effect on public safety, and may require
urgent repair. There may also be a remote possibility that the fracture causes a catastrophic collapse, or a severe consequence. It is necessary to determine which of these consequences is most
realistic and establish sufficient rationale based on experience, engineering judgment, and/or theoretical analysis to exclude those consequences that are not credible scenarios.
While the immediate effect on the structure is primarily what is evaluated (e.g., collapse after
member failure), it is also appropriate to consider longer term consequences. For example, in the
example cited above, if the fracture were to result in a lane closure on a portion of interstate that carries a very high ADTT, the consequence on the traveling public could be high to even severe, though
no concerns regarding the structural performance of the bridge may actually exist. Rather, the resulting impacts on serviceability could be such that a more frequent inspection interval is justified.
There are many cases in which the critical consequence is obvious. There are also many that
require considerable judgment and/or analytical effort to ensure the appropriate CF is selected.
Table B1. General description of the CF categories.
Level

Category

Consequence
on Safety

Consequence on
Serviceability

Low

None

Minor

Moderate

Minor

Moderate

High

Moderate

Major

Severe

Major

Major

Summary Description
Minor effect on serviceability,
no effect on safety
Moderate effect on serviceability,
minor effect on safety
Major effect on serviceability,
moderate effect on safety
Structural collapse/loss of life

Guideline for Evaluating the Consequence Factor 41

In these cases, it is important that the rationale used to support the determination is recorded.
There are many situations in which analysis and/or experience can be used to justify selecting
one scenario over another. However, the level and the type of analysis that is required must be
defined, as well as what constitutes sufficient experience and when it is appropriate to use
experience to justify the categorization of the consequence.
This section describes, through example, situations in which analysis or experience is needed to
justify the selection of an appropriate CF. Since not every situation can be included or foreseen,
the reader must use the information provided and consider it a road map or framework on how
to select the appropriate consequence. The Reliability Assessment Panel (RAP) may use this guidance to develop basic rules or common practices for very common scenarios they anticipate in the
analysis. The RAP should consider existing rules, policies, or common practices within its state
regarding the considerations for identifying structural redundancy and other factors that may
influence the assessment of the consequences. If no rules, policies, or common practices exist, it
may be necessary for the RAP to develop its own basic guidelines before performing consequence
assessments.

B 1.1 Definitions
This section provides definitions for the terms analysis and experience as used in the context
of this document to support the selection of the most appropriate CF.
Analysis: As used herein, refers to the effort put forth using accepted methods of structural
analysis to quantitatively evaluate the outcome of a given event or scenario based on certain initial
conditions. Laboratory and field experimental testing are also acceptable methods that can be used
to demonstrate, quantitatively, the outcome of a given event or scenario. Analysis requirements
may be beyond the scope of most engineering specifications currently used for design and rating,
and special assessments may be required in certain conditions. Hence, the owner and the engineer
must agree upon the level of analysis, loading, material properties, etc. that will be used for the basis
of the analysis. Similarly, any laboratory or field testing must properly simulate or represent in-situ
conditions (i.e., scale of the specimen or test, loading, failure mode, etc.) in order to be considered
acceptable.
Experience: As used herein, refers to the use of previous knowledge alone to qualitatively
evaluate the outcome of a given event based on certain initial conditions. In order to use experience, the user must be able to demonstrate at least the following:
1. The characteristics of the structure being evaluated are identical or sufficiently similar to the
structure for which the RAP has previous documented experience.
2. The result of the damage mode is identical for the bridge(s) used as a reference. For example,
strand fracture as a result of corrosion or impact may be effectively the same. In both cases,
the strand failed.
The information on which the decision is based must be included in the documentation of the
RBI assessment. It may consist of the location, structure type, damage type, reason for selection,
or other rationale and evidence used to form the decision so that a permanent record is available
for future RAPs.

B 2 General Descriptions of Consequence Scenarios


This section provides guidance for the treatment of typical scenarios and situations for each
of the four CF categories. A brief description of each CF category is provided, as well as typical
examples or scenarios for each category. Methods for selecting the appropriate CF for a given

42 Proposed Guideline for Reliability-Based Bridge Inspection Practices

failure scenario are described. This section is intended as guidance for evaluation. Specific situations and scenarios may vary, and the RAP should utilize good engineering judgment supported
with analysis or documented experience where necessary. Local rules, policies, and practices of
the bridge owner should be considered in the assessments.
As stated, when assessing the CF, the immediate and short-term outcomes, or the results of the
failure of an element should be considered. The immediate consequence refers to the structural
integrity and safety of the traveling public when the failure occurs. Considerations include whether
a bridge will remain standing when the damage mode occurs, and whether the traveling public will
remain safe. For example, failure of a load bearing member in a multi-girder redundant bridge is
not expected to cause loss of structural integrity, excessive deflections, or collapse. As a result, the
traveling public is not immediately affected when the failure occurs. Another scenario would
be for a fracture-critical bridge, where the loss of a main member could cause excessive deflection or collapse thereby causing the bridge to be immediately unsafe for the traveling public.
The safety of the structure and the public should be considered for determining the immediate
consequence.
The short-term consequence refers to serviceability concerns and short-term impacts to the
traveling public after a given damage mode occurs. Load posting, repairs, and speed reductions
can be considered serviceability concerns. Lane, sidewalk, or shoulder closures as a result of the
damage mode impact the traveling public and can cause delays. For example, a multi-girder
redundant bridge that experiences the loss of a load bearing member is expected to remain
standing; however, once the failure is discovered, a typical response is to close a lane or shoulder
until the bridge is repaired. Therefore, the traveling public will be affected. The serviceability of
the structure and the impact to the traveling public should be considered when determining the
short-term consequence.
For example, the failure of a member in a multi-girder bridge may be a moderate immediate
consequence because the bridge is expected to remain standing and no excess deflections are
expected to occur. However, if this bridge is located on an interstate located downtown in a
major city, the short-term consequence of the member failure may be high or severe because a
lane closure may be required, which would cause significant traffic delays. Therefore, the CF for
this bridge may be high based upon the short-term consequence.
Tables B2 through B6 provide additional guidance for commonly encountered situations for
bridge decks, typical superstructures, and substructures. These tables provide descriptions of
typical immediate and short-term effects from common damage modes and sample situations.
The tables also include factors the RAP may consider in differentiating CF categories. For example, for the damage mode of spalling in a bridge deck, the CF may be different for a low ADT
bridge than for a high ADT bridge, based on serviceability considerations. The CF may be different for a bridge that crosses a roadway than one that crosses a small stream, based on concerns
regarding debris falling into traffic, etc. These tables are not intended to be comprehensive, but
rather are intended to provide guidance and examples to assist an RAP with developing criteria
for determining the CF for typical damage modes for common bridge designs under analysis.

B 2.1 Low Consequence Event


General Description
Minor effect on serviceability, no effect on safety.

This scenario is the least serious of all the CF categories. The likelihood of structural collapse
resulting from the damage mode is not credible and the effect on the serviceability of the bridge
is minor.

Guideline for Evaluating the Consequence Factor 43


Table B2. Consequence table for deck elements. Assumed damage mode is spalling.
Consequence
for Deck

Low

Descripon
Immediate: Damage to the top of the deck does not present a
safety concern for the traveling public. Falling debris from the
boom of deck does not affect the safety of the public.
Short term: Minimal serviceability concerns may require
maintenance. Lile or no impact to traveling public.
Immediate: Damage to the top of the deck presents a minimal
safety concern to the traveling public. Falling debris from the
boom of deck presents a minimal safety concern.

Moderate

High

Severe

Short term: Moderate serviceability concerns. Speed reducon


may be needed. Traffic is moderately impacted as a result of
lane, shoulder, or sidewalk closure on or under bridge.

Sample Situaons
Bridge carrying low volume and/or
low speed roadway
Bridge with concrete deck over a
non navigable waterway or unused
right of way land
Moderately traveled roadway where
damage would cause minimal delays
Bridge with stay in place forms over
roadway where spalls would not
reach roadway or waterway

Immediate: Damage to the top of the deck presents a moderate


safety concern to the traveling public. Falling debris from the
boom of deck presents a moderate safety concern.

High volume roadway where damage


would cause reducon in posted
speed or potenal for loss of vehicular
control

Short term: Major serviceability concerns. Repairs or speed


reducon may be required. Traffic is greatly impacted as a result
of lane, shoulder, or sidewalk closure on or under bridge.

Bridge without stay in place forms


over heavily traveled waterway or
high volume roadway

Immediate: Damage to the top of the deck presents a major


safety concern to the traveling public. Falling debris presents a
major safety concern. Possible loss of life.
Short term: Potenal for significant traffic delays on or under
the bridge.

Factors to Consider

Bridge over feature where spalling


concrete would result in lane closure,
loss of life, or major traffic delays

Requirements for Selection


In order to select the lowest consequence category, the user must be able to clearly demonstrate
that the consequence of the damage will be benign. Generally speaking, this decision will most
often be based on engineering judgment and experience. Situations in which selection of this
consequence scenario may be appropriate are as follows:
Failure of thin deck overlay.
Spalling in a concrete deck bridge on a low-volume and/or low-speed roadway.
Spalling/corrosion damage in an abutment where the bridge is over a non-navigable waterway

or unused right-of-way land.

B 2.2 Moderate Consequence Event


General Description
Moderate effect on serviceability, minor effect on safety.

This scenario can be characterized by consequences that are classified as moderate in terms
of their outcome. The likelihood of collapse and loss of life is very remote, and there is a minor
effect on the safety of the traveling public.
Requirements for Selection
In order to classify the consequence of a given failure scenario as moderate, the user must
demonstrate that the damage mode will typically result in a serviceability issue. The damage mode

ADT/ADTT
Feature under
Feature carried
Stay in place
forms

44 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table B3. Consequence table for steel superstructure elements. Assumed damage mode is loss of one primary
load carrying member.
Consequence
for Steel
Superstructure

Low

Moderate

Descripon
Immediate: Lile to no impact on structural capacity is
expected based upon structural analysis or documented
experience. Public safety is unaffected.

Rural bridge with low ADT/ADTT

Immediate: Structural capacity is expected to remain


adequate based upon structural analysis or documented
experience.

Bridge over mul use path,


railroad, or lightly traveled
waterway

Short term: Moderate serviceability concerns. Speed


reducon or load posng may be needed. Traffic is
moderately impacted as a result of lane, shoulder, or
sidewalk closure on or under bridge.

Bridge on or over moderate


volume urban roadway or high
volume rural roadway that would
cause moderate delays for drivers

Severe

Short term: Major serviceability concerns. Load posng,


repairs, or speed reducon may be needed. Traffic is greatly
impacted as a result of lane, shoulder, or sidewalk closure
on or under bridge.
Immediate: Structural collapse. Possible loss of life.
Short term: Potenal for significant traffic delays on or
under bridge.

Factors to Consider

Bridge over non navigable


waterway or unused right of way
land

Short term: Minimal serviceability concerns may require


maintenance. Lile or no impact to traveling public.

Immediate: Structural capacity is expected to remain


adequate.
High

Sample Situaons

Bridge with alternate load path(s)


that has an expectaon of
adequate remaining structural
capacity

ADT/ADTT
Feature under
Feature carried
Redundancy
Composite
construcon
Load carrying
capacity/rang

Lane or shoulder closure on or


under roadway that would cause
major delays for drivers
Bridge with high ADT/ADTT that
requires closure

poses no serious threat to the structural integrity of the bridge or to the safety of the public.
Generally, damage that will require repairs that can be addressed in a programmed fashion (i.e.,
non-emergency) would be classified as having a moderate consequence. Member or structural
redundancy should be a consideration, and, in cases where the member is non-redundant, it
may be prudent to classify an event higher in consequence. Situations in which the selection of
this CF may be appropriate are as follows:
Spalling damage in a deck soffit or concrete girder for a bridge over multi-use path, railroad,

or low-volume (<10 ADT) roadway.


Spalling in a concrete deck bridge on a moderate-volume roadway.
Lane or shoulder closure on a bridge carrying a moderate-volume urban roadway or a high-

volume rural roadway that would cause moderate delays for drivers.
Fatigue cracks that require repair but are not the result of primary member stresses, such as

out-of-plane distortion cracks in redundant members


The examples above illustrate some of the element failure scenarios that would typically be
categorized as having moderate consequence. In some cases, failure scenarios that could be
considered more serious can be categorized as having moderate consequences, if analysis or
past experience can be used to better define the outcome of a given scenario. For example,
out-of-plane fatigue cracks are not uncommon in some older steel bridges, and are included in

Guideline for Evaluating the Consequence Factor 45


Table B4. Consequence table for reinforced concrete superstructure elements. Assumed damage mode is
loss of one primary load carrying member.
Consequence
for Concrete
Superstructure

Low

Moderate

High

Severe

Descripon
Immediate: Lile to no impact on structural capacity is
expected based upon structural analysis or
documented experience. Falling debris does not affect
the safety of the public.
Short term: Minimal serviceability concerns may
require maintenance. Lile or no impact to traveling
public.
Immediate: Structural capacity is expected to remain
adequate based upon structural analysis or
documented experience. Falling debris presents a
minimal safety concern to the public.
Short term: Moderate serviceability concerns. Speed
reducon or load posng may be needed. Traffic is
moderately impacted as a result of lane, shoulder, or
sidewalk closure on or under bridge.
Immediate: Structural capacity is expected to remain
adequate. Falling debris presents a moderate safety
concern to the public.
Short term: Major serviceability concerns. Load
posng, repairs, or speed reducon may be needed.
Traffic is greatly impacted as a result of lane, shoulder,
or sidewalk closure on or under bridge.
Immediate: Structural collapse. Falling debris presents
a major safety concern to the public. Possible loss of
life.
Short term: Potenal for significant traffic delays on or
under bridge.

Sample Situaons

Factors to Consider

Bridge over non navigable


waterway or unused right of
way land
Rural bridge with low
ADT/ADTT

Bridge over mul use path,


railroad, or lightly traveled
waterway
Bridge on or over moderate
volume urban roadway or
high volume rural roadway
that would cause moderate
delays for drivers
Bridge with alternate load
path(s) that has an expectaon
of adequate remaining
structural capacity
Lane or shoulder closure on or
under roadway that would
cause major delays for drivers
Bridge over feature where
spalling concrete would result
in lane closure, loss of life, or
significant traffic delays

the examples above. However, other types of fatigue cracks may be more serious. For example,
consider cracking in a single plate of a built-up riveted girder. These types of cracks would normally be expected to be much more serious. They may require categorization as having high or a
severe consequence, if it is assumed that the crack propagates such that the load carrying capacity
of the girder is lost. However, in many cases, riveted built-up members are composed of two
or three cover plates, two angles, and the girder web. If it could be shown by analysis that even
after complete cracking of one of these individual components (e.g., complete cracking of one
of the cover plates) the member still has plenty of reserve capacity, then it might be reasonable
to classify the event as a moderate consequence scenario. The individual making this assessment
would also want to consider overall system redundancy and other factors.
Hence, if analysis can be used to show that a condition that is generally perceived to be more
serious, but is actually not so, then it may be justified to classify the event as having a moderate consequence. Experience may also be utilized to assess if a given failure scenario is a high consequence
event or a moderate consequence event. In cases where a given owner may have had the same or
very similar experience with several other identical or sufficiently similar bridges, the owner may

ADT/ADTT
Feature under
Feature carried
Redundancy
Composite
construcon
Load carrying
capacity/rang

46 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table B5. Consequence table for prestressed concrete superstructure elements. Assumed damage mode is
loss of one primary load carrying member.
Consequence
for Prestressed
Superstructure

Low

Descripon
Immediate: Lile to no impact on structural capacity is
expected based upon structural analysis or documented
experience. Falling debris does not affect the safety of
the public.
Short term: Minimal serviceability concerns may require
maintenance. Lile or no impact to traveling public.

Moderate

High

Immediate: Structural capacity is expected to remain


adequate based upon structural analysis or documented
experience. Falling debris presents a minimal safety
concern to the public.
Short term: Moderate serviceability concerns. Speed
reducon or load posng may be needed. Traffic is
moderately impacted as a result of lane, shoulder, or
sidewalk closure on or under bridge.
Immediate: Structural capacity is expected to remain
adequate. Falling debris presents a moderate safety
concern to the public.
Short term: Major serviceability concerns. Load posng,
repairs, or speed reducon may be needed. Traffic is
greatly impacted as a result of lane, shoulder, or sidewalk
closure on or under bridge.
Immediate: Structural collapse. Falling debris presents a
major safety concern to the public. Possible loss of life.

Severe

Short term: Potenal for significant traffic delays on or


under bridge.

Sample Situaons

Factors to Consider

Bridge over non navigable


waterway or unused right of
way land
Rural bridge with low ADT/ADTT

Bridge over mul use path,


railroad, or lightly traveled
waterway
Bridge on or over moderate
volume urban roadway or high
volume rural roadway that
would cause moderate delays for
drivers

ADT/ADTT
Feature under
Feature carried
Redundancy
Composite
construcon
Load carrying
capacity/rang

Bridge with alternate load


path(s) that has an expectaon
of adequate remaining
structural capacity
Lane or shoulder closure on or
under roadway that would
cause major delays for drivers
Bridge over feature where
spalling concrete may result in
lane closure, loss of life, or
significant traffic delays

be able demonstrate that a lower CF is justifiable. Very high load ratings (e.g., 150% of the minimum required) and redundancy could also be factors to consider when selecting this CF category.
Of course, if experience and judgment are used to determine CF, then sufficient documentation
would need to be available to justify the selection of a given CF.

B 2.3 High Consequence Event


General Description
Major effect on serviceability, moderate effect on safety.

This scenario can be characterized by consequences that are more serious in terms of their
outcome. The likelihood of collapse and loss of life may be more measurable, but still relatively
remote.
Requirements for Selection
The user must be able to demonstrate that the possibility of collapse and loss of life are still
relatively remote when identifying a given failure scenario as having a high consequence. Though
the bridge may require repairs, the outcome would not be catastrophic in nature.

Guideline for Evaluating the Consequence Factor 47


Table B6. Consequence table for substructure elements. Assumed damage mode is spalling.
Consequence
for
Substructure

Low

Descripon
Immediate: Falling debris does not affect the safety of the
public. Structural capacity of the bridge remains adequate.
Short term: Minimal serviceability concerns may require
maintenance. Lile or no impact to traveling public.
Immediate: Falling debris from substructure presents a
minimal safety concern to the public. Structural capacity is
expected to remain adequate based upon structural
analysis or documented experience.

Moderate

Short term: Moderate serviceability concerns. Speed


reducon or load posng may be needed. Traffic is
moderately impacted as a result of lane, shoulder, or
sidewalk closure on or under bridge.
Immediate: Falling debris from substructure presents a
moderate safety concern to the public. Structural capacity
is expected to remain adequate.

High

Short term: Major serviceability concerns. Load posng,


repairs, or speed reducon may be needed. Traffic is greatly
impacted as a result of lane, shoulder, or sidewalk closure
on or under bridge.
Immediate: Structural collapse, bearing area failure, or loss
of load carrying capacity. Falling debris presents a major
safety concern to the public. Possible loss of life.

Severe

Short term: Potenal for significant traffic delays on or


under bridge.

Sample Situaons

Factors to Consider

Bridge over non navigable


waterway or unused right of way
land

Bridge over mul use path,


railroad, or lightly traveled
waterway
ADT/ADTT
Feature under
Load carrying
capacity
Lane or shoulder closure on
roadway that would cause major
delays for drivers

Bridge adjacent to high volume


roadway with insufficient
horizontal clearance where
spalling concrete may result in
lane closure, loss of life, or major
traffic delays
Bearing area failure resulng in
deck misalignment

Examples of high consequence events would include scenarios that require short-term closures
for repairs, lane restrictions that have a major impact on traffic, load postings, or other actions
that majorly affect the public. Situations where the selection of this CF may be appropriate are
as follows:
Failure of a main member in a multi-girder bridge with sufficient load path redundancy.
Spalling damage in a deck soffit or concrete girder for a bridge over a navigable waterway or

a moderate-/high-volume roadway.
Spalling in a concrete deck bridge on a high-volume roadway.
Lane or shoulder closure on or under a roadway that would cause major delays for drivers.
Impact damage on a multi-girder bridge.
Again, using brittle fracture of a girder as an example, consider the response to the fracture of
an exterior girder in a multi-girder bridge. If the girders are spaced relatively closely, a reasonable
strategy would be to place barriers on the bridge to keep traffic off the shoulder and hence, off
the faulted girder. Though one girder out of several was compromised, experience indicates the
remaining girders have sufficient capacity to carry traffic safely.
In the above example, it is important to note the reaction to the fracture was not based on
calculations, but was based entirely upon experience. If the owner performed calculations that

48 Proposed Guideline for Reliability-Based Bridge Inspection Practices

quantifiably showed that the bridge had sufficient reserve capacity in the faulted condition, i.e.,
with one girder fractured, it might be reasonable to identify the event as having a moderate
consequence.
Guidance on such analysis exists in the literature and it can be performed for common bridges
and common bridge types. However, simplified analytical procedures may also suffice. For example, there is considerable discussion regarding redundancy of multi-girder systems, both concrete
and steel, as reported in NCHRP Report 406: Redundancy in Highway Bridge Superstructures (1).
This document provides direction on determining the capacity and the redundancy as a function of span, girder spacing, and the number of loaded lanes using system factors. The research
resulted in the development of system factors that quantify redundancy based on an assessment
of the reliability of the bridge systems, rather than simply the individual bridge members. Using
the recommended system factors may greatly reduce the analytical effort needed in assessing a
bridge. The major conclusion from this research was that bridges designed to AASHTO bridge
specifications generally possess sufficient reserve capacity. In addition, NCHRP Project 12-87,
Fracture-Critical System Analysis for Steel Bridges was underway at the time this report was
prepared and once complete may be of use in performing system analysis.
If experience is used as the reason to justify a reduction from a high consequence to a moderate
consequence, the experience referenced would have to be for a type of structure and a damage mode
outcome that is nearly identical to the one under consideration, as described in section B 1.1. (For
example, corrosion, fatigue, or fracture can all lead to a failed girder. Hence, although the damage modes are different, the outcome is the same.) Therefore, the RAP would have to adequately
document and demonstrate that the cited case(s) are of sufficient similarity. Owners may cite
examples both in their own state and from other states. Another desirable characteristic would
be whether or not the experience with a given response has been observed more than once. For
example, an owner may have experience with a certain type of rolled steel beam bridge and
truck impact. Experience with truck impacts on several similar steel bridges may demonstrate
that for the bridge under consideration, impact to the superstructure would not result in a set
of circumstances that justify identifying the event as having a high consequence. Based on this
experience, it may be appropriate to identify the event as having only moderate consequences.
Another example would be a case in which there is severe spalling at the bearing of a member
in a prestressed, multi-girder bridge that is over a small creek or a flood plain. Hence, there is
no concern regarding spalled concrete hitting someone or something below the bridge (minor
effect on public safety). If calculations could be made to show that if the bearing were to completely fail, there would only be moderate effects on serviceability, then it would be reasonable
to state this is a moderate consequence event. In the absence of detailed calculations and/or
substantial experience regarding the specific scenario, it would be required to be identified as
having a high consequence, based on the criteria discussed.

B 2.4 Severe Consequence Event


General Description
Major effect on serviceability and safety.

This is the most critical CF category and can be characterized by events that, should they
occur, are anticipated to result in catastrophic outcomes. Structural collapse and loss of life are
likely should the failure occur.
Requirements for Selection
Due to the catastrophic nature implied by this consequence scenario, it should not be selected
arbitrarily as a catch-all or just to be conservative. The user must have reasonable justification

Guideline for Evaluating the Consequence Factor 49

that shows that the failure scenario being considered is likely to be consistent with a severe consequence event.
Examples of severe consequence events would include failure of the pin or hanger in a bridge
with a suspended truss span or a two-girder system, or strand fractures in a pre- or posttensioned element that results in a non-composite member falling into a roadway below, such
as what was observed in Washington Township, PA (2). Failure of a pier due to severe corrosion
of the reinforcement or to a lack of reinforcement would also be an example of a severe consequence event. Situations in which the selection of this CF may be appropriate are as follows:
Fracture in a non-redundant steel bridge member.
Failure of a non-composite girder over traffic.
Spalling of a concrete soffit, concrete girder, or concrete abutment over a high-volume road-

way or pedestrian walkway.


Lane or shoulder closure on a major roadway that would cause significant delays for the
traveling public.
Bearing area failure resulting in deck misalignment.
Cases for which there is insufficient experience or where reliable calculations cannot be made
(due to lack of analytical models or data for use in the models) may also be categorized as severe.
Examples would be unique, one-of-a-kind bridges or other structural systems for which the result
of failure associated with a given damage mode is essentially unknown. In such cases, the only
reasonable approach is to conservatively assume and select the worst-case consequence (i.e., a
severe consequence), as the actual outcome cannot be well defined.
A common example of a failure that would result in a severe consequence is primary member
failure in a fracture-critical bridge. Due to the perceived lack of redundancy, fracture of a main
member is assumed to result in a total collapse of a bridge or a portion of a bridge. Though this is a
reasonable conclusion in the absence of more rigorous analysis, the bridges can also be good examples of where more rigorous analysis can be used to show redundancy actually exists. For example, a literature review conducted as part of NCHRP Synthesis 354: Inspection and Management of
Bridges with Fracture-Critical Details, revealed that there were no documented cases of catastrophic
failure for any two-girder bridges or cross girders where fractures had occurred (3). In some of the
failures, an entire girder fractured, but due to inherent redundancy of the unaccounted-for load
paths, such as the deck and lateral system, and overall system behavior, the bridges did not collapse.
In fact, in some cases, there is little perceived deflection in the faulted state.
In light of the above, owners may wish to perform an after-fracture redundancy analysis to
demonstrate that a given bridge possesses sufficient alternate load paths such that the most likely
outcome would have only high consequences. Obviously, the owner must select the appropriate
live load that must be carried in the faulted state for the analysis. Further, consideration should
be given to the fact that the bridge may need to remain in service for some time with the fracture
undetected. For example, if the fracture occurred immediately after a scheduled inspection and
there was little or no evidence that would alert anyone to the condition and to take action (e.g.,
no deflection).
Obviously, there are other damage modes that may result in a severe consequence. For those,
analysis may also be used to demonstrate that the most likely outcome would have only a high
consequence. Downgrading to the less serious consequence scenario is permitted but only
through the use of analysis. Experience alone may not be used to justify downgrading from a
severe consequence to a high consequence, due to the catastrophic outcomes associated with
the more severe scenario. While experience may be used in conjunction with analytical studies
to make a stronger case for downgrading to a lower consequence scenario, experience alone is
not deemed to be sufficient.

50 Proposed Guideline for Reliability-Based Bridge Inspection Practices

B 3Use of Expert Elicitation for Determining


the Consequence Scenario
An expert elicitation of the RAP can be a useful tool for evaluating the appropriate CF for situations that are not well matched to the examples given above, or to establish basic ground rules
for the assessment of common situations. The expert elicitation process can be used to establish
or to build consensus among the RAP and to assist in identifying the most likely outcomes of
damage modes assessed during the reliability analysis. The process should be carefully controlled
and systematic to ensure that the judgments of the RAP are effectively ascertained. The process
involves a few basic, but critical, steps as follows:
1. Statement of the Problem: The RAP should be presented with a clear statement of the problem
and supporting information to allow for expert judgment to be made. Care should be taken
to ensure the problem statement does not contain information that could lead to a biased
decision. The problem statement typically includes data regarding the bridge design, location,
typical traffic patterns, and the failure scenario under consideration.
2. Expert Elicitation: Independently, each member of the RAP is asked, based on his or her judgment, experience, available data, and given the scenario presented, to determine what the most
realistic consequence is resulting from the damage mode under consideration. The expert is
asked to express this as a percentage, with the smallest unit of estimate typically being 10%. The
expert provides a written statement on what factors they considered in making the estimate.
3. Comparison of Results: Once each member of the RAP has rated the situation, the results of
the elicitation are aggregated. Generally, there will be consensus regarding the most critical
consequence. However, in some cases, the most critical choice will not be clear and there will
not be consensus.
4. Identify Consequence Factor: If there is consensus among the panel regarding the appropriate CF,
then the rationale for making the determination is recorded. This rationale should be consistent
with the general guidance herein, or document deviations, changes, and associated rationale.
For cases in which consensus is not reached in the initial elicitation, the experts should discuss
their rankings, their assumptions, and rational for their specific judgments. The members of the
RAP should then be given the opportunity to discuss the various judgments and to revise their
scores based on the discussion. In some cases, additional information may be needed to support
developing a consensus regarding the appropriate CF. For example, analysis may need to be conducted or previous experience documented. If consensus cannot be reached, a potential approach
would be to adopt the most conservative consequence scenario that was included among the
revised scores. Exceptions to the selected CF should also be documented.
When consensus cannot be reached, the RAP may determine that additional analysis is required
to determine the appropriate consequence for a given failure scenario. In some cases, additional
data collection may be required in order to reach a consensus. Regardless of the approach, the
individual RAP should have the flexibility to develop its own methodologies to handle cases for
which there is no consensus. However, at the conclusion of the analysis, the method still must
result in the selection of the most appropriate consequence scenario, based on the guidelines
provided herein and on good engineering judgment.

B 4 References
1.Ghosn, M., Moses, F., NCHRP Report 406: Redundancy in Highway Bridge Superstructures. 1998, TRB,
National Research Council: Washington, D.C.
2.Clay, N., et al., Forensic Examination of a Noncomposite Adjacent Precast Prestressed Concrete Box Beam
Bridge. Journal of Bridge Engineering. 15(4): p. 408418.
3.Connor, R. J., R. Dexter, and H. Mahmoud, NCHRP Synthesis 354: Inspection and Management of Bridges with
Fracture-Critical Details. 2005, Transportation Research Board of the National Academies: Washington, D.C.

APPENDIX C

Guideline for Determining


the Inspection Interval
52
52

C 1 Inspection Intervals
C 1.1 Important or Essential Bridges

51

52 Proposed Guideline for Reliability-Based Bridge Inspection Practices

C 1 Inspection Intervals

Figure C1.Risk
matrix for a typical
highway bridge.

Inspection intervals are determined based on the reliability analysis using a simple four by four
matrix as shown in Figure C1, which illustrates a risk matrix for a typical highway bridge. Engineering judgment is required for establishing the specific divisions applied to the risk matrix; the
divisions are generally applied to ensure that the likelihood of damage remains low during the
interval between inspections, such that there are multiple inspections conducted before there is
a high likelihood of failure occurring. When consequences are relatively high, should the failure
occur, the interval is further reduced to provide an extra margin of safety.
For the risk matrix shown in Figure C1, divisions have been made to separate the bridges
requiring more frequent inspections (Category I) from those requiring less frequent inspections (e.g., Categories III, IV, and V). The inspection interval categories are shown in Table C1.
Bridges with elements falling in Category II require the typical inspection interval of 24 months,
currently used under the NBIS.
The inspection intervals and the divisions on the risk matrix are engineering-based to ensure
a high margin of safety and that multiple periodic inspections take place before the likelihood of
failure becomes high. In other words, the intervals are determined such that the likelihood of
failure remains low, and the intervals are further reduced as consequences increase to provide
additional levels of safety. For example, recall that the RAP assessment of the likelihood of a
damage mode resulting in a failure (as defined in Section 2.1) is based on a 72-month timeframe. For a given element, if there is low likelihood of a failure (OF = 2), and the consequence
of that failure is moderate (CF = 2), the inspection interval of 72 months (Class IV) is identified
on the matrix. This is justified because the analysis has indicated that there is a low likelihood of
failure, and even if the failure occurs, there will be only a moderate effect on the serviceability
of the bridge. However, if the consequence of the failure were high, then the inspection interval is reduced to 48 months (Class III) and 24 months (Class II) if the consequence is severe.
Alternatively, if the likelihood of failure is moderate (OF = 3) over 72 months, the maximum
inspection interval is less than 72 months, regardless of the consequence; 48 months if the consequence were only low (benign) (CF = 1) or moderate (CF = 2) and 24 months if the consequence were high (CF = 3). Similarly, if the likelihood of failure were remote over the 72-month
timeframe, it may be justified to have a maximum interval of more the 72 months, particularly
if the consequences are assessed to be benign (CF = 1). As the consequences increase, this
interval is reduced.

C 1.1 Important or Essential Bridges


As noted, the divisions on the risk matrix require engineering judgment to determine which
inspection intervals are acceptable and necessary. For certain bridges, for example, essential
bridges along key transportation routes, an owner may wish to provide an additional margin of
reliability. Under these circumstances, the divisions on the risk matrix may be adjusted down

Table C1. Maximum inspection interval


categories.
Category

Maximum Interval

I
II
III
IV
V

12 months or less
24 months
48 months
72 months
96 months

Guideline for Determining the Inspection Interval 53

Figure C2. Risk matrix


that may be applied to
essential bridges.

and toward the lower left corner of the matrix. For example, Figure C2 illustrates a risk matrix an
owner could apply to bridges for which an additional measure of reliability is desired. This may
be due to the importance of the bridge to the effectiveness of the transportation system overall,
and/or because the bridge serves essential purposes. Criteria for identifying these essential or
important bridges should be developed by the bridge owner, but would typically consider such
factors as ADT, functional classification of the route, and importance to local transportation
functions. Owners may already have criteria for identifying essential or important bridges for
which added measures of reliability are desired.

APPENDIX D

Inspection Technologies

55
55

56

54

D 1 Introduction
D 1.1 NDE Method Technical Readiness Levels and Costs

D 2 Inspection Methods and Technologies

Inspection Technologies 55

D 1 Introduction
This appendix provides general guidance for the inspection methods to be utilized in a
risk-based inspection (RBI) practice. The section includes a description of nondestructive
evaluation (NDE) technologys technical readiness and relative costs to assist decision makers in determining appropriate and practical technologies for the detection and evaluation
of typical damage modes and deterioration mechanisms in highway bridges. This section
also includes tables that indicate the relative reliability of different inspection methods and
NDE technologies to assist decision makers regarding the application and effectiveness of
the technologies.

D 1.1 NDE Method Technical Readiness Levels and Costs


This section provide general guidance on the technical readiness levels (TRLs) and costs of the
most common NDE technologies that may be applied for damage detection in highway bridges.
Technologies have been evaluated on relative scales using expert judgment and experience.
Table D1 indicates the scale used to assess the TRL of the technologies. This scale is intended to
assist engineers in understanding the practicality and the availability of NDE technologies, and
to discriminate between those techniques that are readily available and well proven, from those
that may be more experimental in nature. The scales provide a five-level discrimination that
indicates if an NDE technology is experimental in nature, or if it is a widely available and widely
implemented technology.
NDE technologies are rated according to the cost scales shown in Table D2. These scales are
intended to provide engineers with general information regarding the relative costs of implementing NDE technologies for bridge inspection. Relative costs are based on a typical, multigirder highway bridge approximately 150 ft in length.
The TRL and costs for NDE technologies are shown in Table D3.

Table D1. TRLs for NDE technologies.

Description

Technical Readiness Level


TRL
No.

Fundamental Research: basic research in the


laboratory

In Development: laboratory equipment, starting


field testing and experimental applications, proof
of concept testing

Application Development: Applications for the


technology are being developed, commercially
available research equipment, field testing is
experimental/developmental, initial assessments of
effectiveness in the field, reliability unknown

Controlled Implementation: Commercially


available equipment and service, application by
specialist/consultant, (certification may be
available), assessments of reliability/effectiveness
are ongoing
Widespread Implementation: Certification
available, widely used, commercially available
equipment, commonly available, application by
suitably trained technician, generally accepted
reliability/effectiveness

Examples
Fundamental sensor research,
nano-sensors, laser-induced
breakdown spectroscopy (LIBS)
In-situ corrosion sensors, positron
annihilation, backscatter x-ray,
thermal crack detection
Electromagnetic-acoustic
transducer (EMAT) sensors,
ultrasonic stress measurement,
magnetic flux leakage for
embedded strands thermal crack
detection

Ground penetrating radar (GPR),


radiography, impulse response,
phased array ultrasonics, infared
thermography (IR)

Ultrasonic pulse velocity (UPV),


dye penetrant, eddy current,
magnetic particle, covermeters,
half cell

56 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table D2. Relative cost scales for NDE technologies.
Cost Scales

Symbol

Description
Low cost, state forces, or $100s of dollars to
apply/bridge
Moderate Cost, $1,000$10,000 typical costs/bridge
High cost, >$10,000 to apply

Examples
Dye penetrant, magnetic
particle, impact echo,
ultrasonic thickness,
thermography
GPR, ultrasonic crack
detection, impact echo
Health monitoring, x-ray
diffraction, radiography

Table D3. TRL and cost for typical NDE technologies.


Code
MP
PT
UT
UT-T
ET
AE
IR
GPR
UPV

Name
Magnetic particle
testing
Dye penetrant
testing

TRL

Ultrasonic testing
Ultrasonic
thickness gage
Eddy current
testing
Acoustic emission
Infrared
thermography
Ground
penetrating radar
Ultrasonic pulse
velocity

Cost

Material

Primary Usage

Steel

Surface-breaking cracks in steel

Steel

Surface-breaking cracks in steel

Steel

Surface and subsurface cracks in steel, volumetric defects

Steel

Plate thickness, section loss

Steel

Surface-breaking cracks in steel

Steel

Monitoring growth of fatigue cracks

Concrete

Concrete

Concrete

Subsurface delaminations in concrete


Detecting damage in concrete associated with corrosion, rebar depth,
locating embedded metal objects
Deterioration of concrete, concrete moduli/strength, subsurface voids,
cracks

IE

Impact echo

4/5

Concrete

Delaminations in concrete, deterioration of concrete, subsurface voids

CD

Chain drag

Concrete

Delaminations in concrete

HC

Half-cell potential

Concrete

Corrosion potential

Concrete

Internal voids, loss of section/fracture in embedded steel

Concrete

Delaminations, deterioration of concrete

Concrete

Cracking and deterioration in concrete, delaminations

Concrete

Loss of section for embedded steel element (prestressing strand, rebar)

RT
S
SAW
MFL

Radiographic
testing
Sounding
Surface acoustic
wave
Magnetic flux
leakage

D 2 Inspection Methods and Technologies


The tables included in this section (Tables D4 through D9) qualitatively describe the reliability
and effectiveness of NDE technologies and inspection methods including routine inspection and
hands-on inspections. In making the assessments of reliability and effectiveness, it was assumed
that a routine inspection was conducted without hands-on access to the bridge element. The
reliability assessment indicated in Tables D5 through D9 is intended to provide general guidance on effective inspection methods for detecting and evaluating certain damage modes and
deterioration mechanisms. Key monitoring or sampling methods that provide tools for assessing
the likelihood of corrosion damage developing in concrete have been included. The key to the
symbolic guide is shown in Table D4. Methods that are low reliability typical do not provide
effective detection or assessment, and are not recommended for the damage mode or deterioration mechanism indicated.

Inspection Technologies 57
Table D4. Symbolic guide to
inspection method reliability
and effectiveness.
Key
Low
Moderate - low
Moderate - high
High

Table D5. Inspection methods for bare concrete decks.


Damage Mode
or Mechanism

Routine
Visual

Hands-On
Visual

Sounding1

IR

GPR

Impact
Echo

Chain
Drag

Spalling/patches
Delamination
(dry)
Deck cracking
(distributed)
Corrosion
damage
Freeze-thaw/
pulverized/
cracks
Delamination
in soffit2

NA

ASR

Half
Cell

Chloride
Ion
Content

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

Active corrosion/
corrosion
potential
1

Based on FHWA visual inspection study results.

NDE technologies applied to the soffit surface.

Table D6. Inspection methods for concrete decks with overlays.


Half
Cell

Chloride
Ion
Content

Spalling/patches

NA

NA

Delamination

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

Damage Mode
or Mechanism

Routine
Visual

Hands-On
Visual

Debonding/
overlay
delamination
Corrosion
damage
Freeze-thaw/
pulverized/
cracks
Delamination
in soffit1
ASR
Active corrosion/
corrosion
potential
1

NDE technologies applied to the soffit surface.

Sounding

IR

GPR

Impact
Echo

Chain
Drag

NA

58 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Table D7. Concrete deck overlays.


Damage
Mode or
Mechanism

Routine
Visual

Hands-On
Visual

Sounding

IR

GPR

Impact
Echo

Chain
Drag

SAW

Spalling/patches
Delamination/
debonding
Overlay
cracking

Table D8. Inspection methods for steel members.


Damage
Mode or
Mechanism
Fatigue cracks
primary
stress
Out of plane
distortion
cracking

Routine
Visual

Hands-on
Visual1

PT2

MT2

UT2

UT-T

ET2

NA

Section loss
Coatings
failure
Steel pins
pack rust
Cracks in
steel pins

NA

NA

NA

NA

NA

NA

NA

NA

NA

NA

Assumes inspectors have been adequately trained.

Assumes Level II certification; Level III procedure development.

NA
NA

NA

Table D9. Inspection methods for open prestressed girders.


Damage Mode
or Mechanism

Routine
Visual

Hands-on
Visual

IR

GPR

IE

MFL

RT

UPV

Chloride
Ion
Content

Spalling/patches

NA

Delamination

NA

Strand corrosion

NA

Freeze-thaw/
pulverized/
cracks
Delamination
in soffit1
ASR
Active corrosion/
corrosion
potential
1

Sounding

NDE technologies applied to the soffit surface.

NA
NA
NA

APPENDIX E

Attribute Index and Commentary

61 Introduction
61
62
62
62
63
63
64
65
65
65
66
66
67
67
67
67
69
69
69
71
72
73
73
73
75
76
76
76
77
78
78
78
79
79
79

Scoring Scheme
Screening Attributes
S.1 Current Condition Rating
S.2 Fire Damage
S.3 Susceptible to Collision
S.4 Flexural Cracking
S.5 Shear Cracking
S.6 Longitudinal Cracking in Prestressed Elements
S.7 Active Fatigue Cracks Due to Primary Stress Ranges
S.8 Details Susceptible to Constraint-Induced Fracture (CIF)
S.9 Significant Level of Active Corrosion or Section Loss
S.10 Design Features
Design Attributes
D.1 Joint Type
D.2 Load Posting
D.3 Minimum Vertical Clearance
D.4 Poor Deck Drainage and Ponding
D.5 Use of Open Decking
D.6 Year of Construction
D.7 Application of Protective Systems
D.8 Concrete Mix Design
D.9 Deck Form Type
D.10 Deck Overlays
D.11 Minimum Concrete Cover
D.12 Reinforcement Type
D.13 Built-Up Member
D.14 Constructed of High Performance Steel
D.15 Constructed of Weathering Steel
D.16 Element Connection Type
D.17 Worst Fatigue Detail Category
D.18 Skew
D.19 Presence of Cold Joints
D.20 Construction Techniques and Specifications
D.21 Footing Type
D.22 Subsurface Soil Condition

59

60 Proposed Guideline for Reliability-Based Bridge Inspection Practices

80
80
81
81
82
82
83
83
84
84
84
85
86
86
86
87
87
88
88
89
89
90
90
91
92
92
93
93
94
94
94
95

Loading Attributes
L.1 ADTT
L.2 Dynamic Loading from Riding Surface
L.3 Exposure Environment
L.4 Likelihood of Overload
L.5 Rate of De-icing Chemical Application
L.6 Subjected to Overspray
L.7 Remaining Fatigue Life
L.8 Overtopping/High Water
Condition Attributes
C.1 Current Condition Rating
C.2 Current Element Condition State
C.3 Evidence of Rotation or Settlement
C.4 Joint Condition
C.5 Maintenance Cycle
C.6 Previously Impacted
C.7 Quality of Deck Drainage System
C.8 Corrosion-Induced Cracking
C.9 General Cracking
C.10 Delaminations
C.11 Presence of Repaired Areas
C.12 Presence of Spalling
C.13 Efflorescence/Staining
C.14 Flexural Cracking
C.15 Shear Cracking
C.16 Longitudinal Cracking in Prestressed Elements
C.17 Coating Condition
C.18 Condition of Fatigue Cracks
C.19 Presence of Fatigue Cracks due to Secondary or Out of Plane Stress
C.20 Non-Fatigue-Related Cracks or Defects
C.21 Presence of Active Corrosion
C.22 Presence of Debris

95 References

Attribute Index and Commentary 61

Introduction
This section includes suggested attributes for the reliability assessment of bridges. Users can
select attributes from this listing. It is also recommended that users develop additional attributes
that meet the needs of their individual agencies. This commentary is organized into four sections: Screening, Design, Loading, and Condition. The Screening section describes attributes that
may be used to quickly identify bridges that should not be included in a particular analysis, either
because they already have significant damage or they have attributes that are outside the scope
of the analysis being developed. In many cases, these attributes may require engineering analysis
beyond that which is typically conducted during a reliability assessment using this Guideline.
Screening attributes are typically attributes that:
Make the likelihood of failure very high.
Make the likelihood of failure unusually uncertain.
Identify a bridge with different anticipated deterioration patterns than other bridges in a group.

Design attributes are characteristics of a bridge element that are part of the elements design.
Design attributes are frequently intrinsic characteristics of the element that do not change over
time, such as the amount of concrete cover or material of construction [concrete, high performance concrete (HPC), etc.]. In some cases, preservation or maintenance activities that contribute to the durability of the bridge element may be a design attribute, such as the use of
penetrating sealers as a preservation strategy.
Loading attributes are characteristics that describe the loads applied to the bridge element.
These may include structural loading, traffic loading, or environmental loading. Environmental
loading may be described in macro terms, such as the general environment in which the bridge
is located, or on a local basis, such as the rate of de-icing chemical application on a bridge deck.
Loading attributes describe key loading characteristics that contribute to the damage modes and
deterioration mechanisms under consideration.
Condition attributes describe the relevant bridge element conditions that are indicative of its
future reliability. These can include its current element or component level rating, or may be a
specific condition that will affect the durability of the element. For example, if the deterioration
mechanism under consideration is corrosion at the bearing areas, the condition of the bridge joint
may be a key attribute in determining the likelihood that corrosion will occur in the bearing area.
The listing of attributes included here is not intended to be comprehensive or mandatory.
Users should consider adding attributes that are important to their specific inventory. Users are
encouraged to document the rationale for including additional attributes in the reliability assessment, along with an appropriate scoring scheme. Users may also wish to omit certain attributes
if they are not relevant to their inventory or do not contribute to the reliability and durability of
bridges within their inventory. The suggested weightings are also exemplary in nature and may
need to be adjusted to meet the needs of a particular bridge inventory.

Scoring Scheme
Attributes are assigned points based on the importance or contribution of the attribute in
terms of the durability and the reliability of the element being assessed. In general, the scoring
scheme utilizes a three-stage assessment of the importance of the attribute as shown in Table E1.
The Ranking Descriptor is intended to provide some verbal description of the weight associated
with each score. As shown, three relative course levels are presented: Low, Moderate, and High.
The RAP may wish to modify the suggested scoring for a given attribute, based on local conditions, past experience, and previous performance within its bridge inventory and operational

62 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table E1. Suggested rank
scoring for attributes.
Ranking Descriptor

Total Points

High

20

Moderate

15

Low

10

environment. The scoring scheme should effectively develop sound engineering rationale to
support risk-based inspection practices.

Screening Attributes
S.1 Current Condition Rating
Reason(s) for Attribute. The current condition rating characterizes the overall condition of the component being rated according to the NBIS rating scale. Bridge components
that have condition ratings of 4 or less have been rated to be in poor condition. In some cases,
these components may already be on a reduced (12 month or less) inspection frequency.
Users may wish to use this criterion to screen bridges that are already in poor condition and,
as a result, require more in-depth analysis to identify their inspection needs. Users could also
assign the OF of high without further assessment, since the component is already in poor
condition.
For element-level inspection approaches, National Bridge Elements (NBEs) or Bridge Management Elements (BMEs) could be utilized within the screening criteria, as appropriate for
specific bridge inventories and inspection practices. Generally, elements indicated with condition states of 4 would be appropriate for consideration as a screening tool for elements selected
to match the needs and practices within the specific bridge inventory.
Assessment Procedure. This screening attribute is scored based on whether the current
condition rating is 4 or less or greater than 4. The current condition rating from the most recent
inspection report should be used. If using an element-level approach, the RAP should identify
appropriate elements and condition states for screening.
Current condition rating is less than
or equal to 4

Component is screened from general


reliability assessment

Current condition rating is greater than 4

Continue with procedure

S.2 Fire Damage


Reason(s) for Attribute. Incidences of fire on or below a highway bridge are not uncommon. This type of damage is most frequently caused by vehicular accidents that result in fire,
but secondary causes such as vandalism, terrorism, or other damage initiators should not be
discounted. If fire does occur on or below a bridge, an appropriate follow-up assessment should
be conducted to determine how the fire has affected the load carrying capacity and the durability characteristics of the main structural members and the deck. This assessment is typically
performed during a damage inspection immediately following the incident.
Damage to bridge components resulting from a fire is either immediately apparent during
the damage inspection, or may manifest within the first 12- to 24-month interval following the

Attribute Index and Commentary 63

fire. Based on this observation, bridges that have experienced a fire may be screened from the
reliability assessment until an inspection, which has been conducted approximately 12 months
or more after the fire, confirms that the fire has not affected the typical durability characteristics
of the bridge components. The purpose of this screening is to ensure that damage from the fire
has not manifested after the damage inspection.
Assessment Procedure. This attribute is scored based only on the occurrence of a fire on
or below the structure being assessed. It is assumed that an appropriate assessment immediately
following the fire incident (i.e., damage inspection) has been performed.
Fire incident has occurred and an inspection
12 months after the fire has not occurred

Bridge is not eligible for reliability


assessment until inspection confirms
that the bridge is undamaged

There have been no incidence of fire on or


below the bridge, or inspections conducted
approximately 12 months or more an
after the fire have confirmed that the
bridge is undamaged

Continue with procedure

S.3 Susceptible to Collision


Reason(s) for Attribute. This screening attribute can be used to screen an inventory or
a family of bridges to identify those bridges with specific vulnerabilities to random or nearrandom damage from collision. This attribute is intended to apply to a limited number of
bridges for which the risk of collision is unusually high or special. Simply because a bridge could
be subjected to impact does not mean the likelihood of impact is high, and, in fact, it could actually be quite remote. However, there are some structures that have been impacted many times
in the past, where a channel or a roadway is particularly difficult to navigate, vertical clearance is
inadequate, etc. that are much more likely to be struck. Examples include collisions from barges,
debris, or heavy trucks. This attribute would typically be used to screen specific bridges that have
an unusual or a unique risk of collision damage than a larger group or family of bridges which
do not. In such cases, individual reliability analysis may be required.
Assessment Procedure. This screening attribute should be assessed based on sound engineering judgment and is intended to screen bridges with unusual or special collision risks from
an assessment of a group of bridges that do not.
Highly susceptible to collisions

Requires specialized assessment and/or mitigation

Structure is not susceptible to collisions

Continue with procedure

S.4 Flexural Cracking


Reason(s) for Attribute. When the primary load-bearing members in a concrete bridge
exhibit flexural cracking, it may indicate that the members were either inadequately designed
for the required loading, that overloads have occurred, or that deterioration has occurred that
has reduced the load-bearing capacity of the members. In any case, large flexural cracks can be
indicative of an inadequate load-bearing capacity that may require an engineering analysis in
order to determine the cause of the cracking and the resulting effect on the load capacity of the
structure. As a result, bridges exhibiting moderate to severe flexural cracking should be screened
from the general reliability assessment unless appropriate engineering analysis indicates that the
cracking is benign or corrective repairs have been made.

64 Proposed Guideline for Reliability-Based Bridge Inspection Practices

The effects on the strength and the durability of a prestressed element due to flexural cracking
are generally more significant than for a reinforced concrete element.
Assessment Procedure. Flexural cracks will typically present themselves with a vertical orientation either near the bottom flange at mid-span or near the top flange over intermediate
supports, if the member is continuous.
Engineering judgment must be exercised in determining whether any present flexural cracking is moderate to severe. Crack widths in reinforced concrete bridges exceeding 0.006 inches to
0.012 inches reflect the lower bound of moderate cracking. The American Concrete Institute
Committee Report 224R-01 (1) presents guidance on what could be considered reasonable or
tolerable crack widths at the tensile face of reinforced concrete structures for typical conditions.
These values range from 0.006 inches for marine or seawater spray environments to 0.007 inches
for structures exposed to de-icing chemicals, to 0.012 inches for structures in humid, moist environments. In prestressed concrete bridge structural elements, tolerable crack width criteria have
been adopted in the Precast/Prestressed Concrete Institute (PCI) Manual for the Evaluation and
Repair of Precast, Prestressed Concrete Bridge Products (MNL-37-06). The PCI Bridge Committee
recommends that flexural cracks greater in width than 0.006 inches should be evaluated to affirm
adequate design and performance.
Presence of moderate to severe flexural
cracking in reinforced or prestressed concrete
bridge elements

Assess individually to determine source,


extent, and effect of cracking

Flexural cracking is not present, or it has


been determined to be benign or repaired

Continue with procedure

S.5 Shear Cracking


Reason(s) for Attribute. If the primary load-bearing members in a reinforced or a prestressed concrete bridge exhibit shear cracking, it may indicate that the members were either
inadequately designed for the required loading, an overload has occurred, or that deterioration has occurred that has reduced the load-bearing capacity of the members. In any case,
shear cracks can be indicative of an inadequate load-bearing capacity requiring an engineering analysis in order to determine the cause of the cracking and the resulting effect on
the load capacity of the structure. As a result, bridges exhibiting cracking attributable to a
deficiency in shear strength should be screened from the reliability assessment unless appropriate engineering analysis indicates that the cracking is benign or corrective repairs have
been made.
Assessment Procedure. Engineering judgment must be exercised in determining whether
any present shear cracking is attributed to a shear strength deficiency. Shear cracks will typically
present themselves with a roughly 45 degree diagonal orientation for conventionally reinforced
concrete and down to roughly 30 degrees for prestressed elements, and will generally radiate
toward the mid-span of the member. The ends of the member and any sections located over
piers should be checked for this type of cracking.
Presence of unresolved shear cracking

Assess individually to determine source


and extent of cracking

Shear cracking is not present or it has been


determined to be benign

Continue with procedure

Attribute Index and Commentary 65

S.6 Longitudinal Cracking in Prestressed Elements


Reason(s) for Attribute. This attribute is for the assessment of prestressed bridge elements.
Longitudinal cracking in prestressed elements can be indicative of corrosion or fracture of the
embedded prestressing strands. As a result, prestressed elements with reported longitudinal
cracking should be individually assessed to determine the source of the cracking and the condition of the prestressing strands.
Assessment Procedure. This attribute is assessed based on data in the inspection report and
engineering judgment. If longitudinal cracking is reported, further assessment may be required.
Significant longitudinal cracking is present

Assess individually to determine source and


extent of cracking and condition of strand

No significant longitudinal cracking

Continue with procedure

S.7 Active Fatigue Cracks Due to Primary Stress Ranges


Reason(s) for Attribute. Active fatigue cracks in steel bridge elements due to primary
stresses can propagate quickly and potentially lead to a fracture in the element. These cracks are
distinguished from distortion cracks or out-of-plane fatigue cracks, which are more commonly
observed, but generally less critical.
Assessment Procedure. If any active fatigue cracks due to primary stresses are found in the
element, it is strongly recommended that the element be retrofitted before continuing with this
procedure. It is noted that a stable fatigue crack can potentially propagate to brittle fracture
depending on the toughness of the material, the total applied stress, and the temperature. A
fatigue crack can be considered not active if previous inspection reports show that the crack
has not grown over a set period of time (e.g., the longest inspection interval plus 1 year). Primary stresses are those stresses (i.e., stress ranges) that are readily calculated using traditional
mechanics principles (e.g., MC/I or P/A) and are typically obtained during design or rating.
Active fatigue crack(s) due to primary
stresses present

Retrofit before continuing

No active fatigue crack(s) due to primary


stresses present

Continue with procedure

S.8 Details Susceptible to Constraint-Induced Fracture (CIF)


Reason(s) for Attribute. Details that are susceptible to CIF can lead to brittle fracture in
the absence of any observable cracking. An example of this is the failure of the Hoan Bridge in
December 2000 in Milwaukee, WI (2). The bridge had been in service for approximately 25 years
before two of the three girders experienced full-depth fractures and the third girder had a crack
that arrested in the flange. Inspection is not a valid method to prevent these types of failures from
occurring (the Hoan Bridge was inspected a few days prior to the failure). Hence, the attribute
is included as a screening criterion.
Assessment Procedure. Details susceptible to CIF have a much higher probability of fracture failure than other types of details. It is recommended that CIF details be retrofitted or
examined more closely before continuing with this process.
Structure contains details susceptible to CIF

Retrofit before continuing

Structure does not contain details


susceptible to CIF

Continue with procedure

66 Proposed Guideline for Reliability-Based Bridge Inspection Practices

S.9 Significant Level of Active Corrosion or Section Loss


Reason(s) for Attribute. This attribute is intended to be used to screen bridges that have
a significant level of existing or active corrosion sites that make the likelihood of severe corrosion damage relatively high. A significant amount of active corrosion and/or section loss in an
element increases the probability of severe corrosion damage developing in the near future. As
a result, individual engineering assessments may be required to effectively assess the reliability
characteristics for the element. Significant section loss would normally be visible for steel structural members.
Assessment Procedure. If a significant amount of active corrosion with section loss is
found on a steel element it is recommended that the element be repaired before continuing with
this process. Engineering judgment must be used to determine what is defined as a significant
amount of active corrosion with section loss and assess its effects. Previous inspection reports
and engineering judgment must also be used to determine whether or not the corrosion is active.
Corrosion damage in steel elements that is inactive is explicitly distinguished from corrosion
that is active. For example, section loss on a girder web that was the result of a leaking expansion
joint that was corrected (the joint was replaced and the girder was repainted), could be classified
as inactive corrosion if the expansion joint repair eliminates the vulnerability to corrosion. It is
assumed that the owner has either determined that the existing section loss is insignificant or has
taken it into account in the rating procedures and load posting, if needed, is in place.
Significant level of active corrosion
and section loss

Repair before proceeding

Active corrosion or section loss is not


significant or has been repaired

Continue with procedure

S.10 Design Features


Reason(s) for Attribute. This attribute is intended to be used to screen bridges that have
unusual or unique design features that make the likelihood of serious damage either usually high
or unusually uncertain, relative to other bridge in the same family, or identify bridges with different anticipated deterioration patterns than other bridges in a group or family. This attribute
can be used to subdivide a family of bridges into two or more groups with similarly anticipated
deterioration patterns, based on specific design features that are not common to each sub-group.
Design features for use as screening items should be identified by the RAP. Two examples below
are provided to illustrate the way in which this attribute might be used.
Bridges with pin and hanger connections: Pin and hanger connections generally have a history of
presenting maintenance challenges. As such, it may be desirable to screen a bridge that includes this
particular type of connection from a larger family, such as a family of steel multi-girder bridges.
Jointless bridges: Jointless bridges are typically less susceptible to corrosion-related damage
associated with leaking joints in the bearing areas. As such, the deterioration patterns may differ
from other bridges of similar materials and general overall design.
Assessment Procedure. Unique or unusual design features should be identified through
review of bridge plans or other documentation describing the design features of a bridge.
Bridge has unique or unusual design feature

Screen

Bridge does not have unique or unusual


design features

Proceed

Attribute Index and Commentary 67

Design Attributes
D.1 Joint Type
Reason(s) for Attribute. Bridge joint types can be categorized as either closed systems or
open systems. Compared to open joint systems, closed joint systems provide for higher durability
based on the way their designs shield the inner workings of the joint from dirt and debris. This,
in turn, increases the amount of time before a joint begins to leak onto other bridge components.
The presence of open-type deck joints increases the probability of chloride-contaminated water
leaking onto bridge elements below the deck, thus increasing the likelihood of corrosion-related
damage.
Assessment Procedure. This attribute is rated based on the presence of open joints.
Open joint system

10 points

Closed joint system

0 points

D.2 Load Posting


Reason(s) for Attribute. The presence of a load posting typically indicates that the given
bridge was either not designed to carry modern loading or that the bridge has become damaged
and its structural capacity has been reduced. A structure of this type may be more likely to experience damage from heavy traffic and dynamic loading. This attribute is intended to consider
the contribution of high and possibly even excessive loads on accelerating damage generally
for a given bridge or a family of bridges. Engineering judgment is necessary to evaluate if this
attribute is applicable. Considerations include the likelihood of the applied loading being higher
than (i.e., illegal) or near the load posting. In some cases, traffic patterns are such that the fact
that the bridge is load posted will not affect the rate of damage accumulation on the bridge. For
example, a bridge is load posted for the states legal truck load, but is located on a parkway where
trucks are prohibited.
Assessment Procedure. This attribute is scored based only on whether or not a bridge has
been load posted; the level of the rating does not need to be considered. This assessment should
consider if the load posting has a significant effect on the durability of the bridge.
Structure is load posted

20 points

Structure is not load posted

0 points

D.3 Minimum Vertical Clearance


Reason(s) for Attribute. This attribute is intended to consider the likelihood that a bridge
may be impacted by an over-height vehicle and damaged such that the deterioration rate of
the superstructure elements may be increased. For concrete bridges, impacts may damage the
embedded reinforcement or the prestressing strands, or damage the typical concrete cover
exposing the steel to the environment. For steel bridges, impacts can deform members and
damage coating systems in the areas of the impact. Impact damage that affects the structural
capacity of the bridge requires a damage inspection and an assessment beyond the scope of a
typical reliability assessment. Users may wish to use this attribute to include the potential for
increased deterioration rates for bridges that experience frequent impact damage.
The bridge superstructures minimum vertical clearance influences on how often it will be
impacted. A bridge with a lower vertical clearance will be more likely to experience impact
damage than a bridge with higher vertical clearance. The likelihood of being hit may also

68 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table E2. FHWA coding guide minimum vertical underclearance
provisions.

Underclearance
Code
9
8
7
6
5
4
3
2
1
0

Minimum Vertical Underclearance


Functional Class
Other
Major and
Interstate and
Principal and
Minor
Other Freeway
Minor
Collectors and
Arterials
Locals
>17 ft
>16.5 ft
>16.5 ft
17 ft
16.5 ft
16.5 ft
16.75 ft
15.5 ft
15.5 ft
16.5 ft
14.5 ft
14.5 ft
15.75 ft
14.25 ft
14.25 ft
15 ft
14 ft
14 ft
Rating <4 and requiring corrective action
Rating <4 and requiring replacement
No value indicated
Bridge closed

Railroad
>23 ft
23 ft
22.5 ft
22 ft
21 ft
20 ft

depend on the traffic composition of the roadway below, such as the average daily truck traffic (ADTT).
This attribute is generally based on the total vertical clearance between the bottom of the
girders and the riding surface of the roadway below. The functional classification of the roadway
below the bridge may also be a consideration. NBIS data fields record the vertical clearance and
the functional classification of the route passing under the bridge, and are rated using the model
provided in the coding guide (3), which is provided in Table E2.
Assessment Procedure. This attribute should be scored based on appropriate measurements or on the information stored in the bridge file. The suggested scoring models shown
below consider only the vertical clearance of the bridges. Users may wish to consider the functional classification or the typical traffic pattens below the bridge in their assessment. In the
scoring models shown, increased importance is given to over height clearances for prestressed
concrete bridges relative to steel and conventionally reinforced bridges. This is due to the potential for strand corrosion when the concrete cover is damaged by impact, and the increased rate
of deterioration for strands relative to mild steel.
Prestressed Concrete Girders
Vertical clearance is 15 feet or less

20 points

Vertical clearance is between 15 feet and 16 feet

15 points

Vertical clearance is between 16 feet and 17 feet

10 points

Vertical clearance is greater than 17 feet or no


under traffic present

0 points

Steel and Concrete Girders


Vertical clearance is 14 feet or less

15 points

Vertical clearance is between 14 feet and 15 feet

12 points

Vertical clearance is between 15 feet and 17 feet

7 points

Vertical clearance is greater than 17 feet or no


under traffic present

0 points

Attribute Index and Commentary 69

D.4 Poor Deck Drainage and Ponding


Reason(s) for Attribute. This attribute is intended to consider the adverse effects of poorly
designed deck drainage systems and the possibility of ponding on the deck surface, as well as for
inadequate provisions for preventing scuppers and drains from splashing de-icing chemicals
onto the superstructure below. Ineffective deck drainage increases the likelihood of bridge elements developing corrosion related damage. This results from drainage onto the superstructure
and the substructure elements. Both concrete and steel elements will have an increased susceptibility to corrosion damage when exposed to prolonged periods of wetness and/or frequent wetdry cycles. The presence of chlorides from de-icing chemicals applied to the deck also increases
the likelihood of corrosion damage to these elements.
This attribute can also be used to characterize decks with ponding or with drain diversion
issues. When water is allowed to sit on the surface of the deck, there is an increase in the likelihood that corrosion of the reinforcing steel will initiate and damage will propagate. Water and
chlorides are more likely to penetrate to the level of the reinforcement when periods of wetness
are prolonged and chloride concentrations at the surface are high.
Assessment Procedure. This attribute is scored based on the drainage design of the bridge
and any known ponding or drainage issues, as noted in the inspection report. Drainage systems
which normally allow water to run off onto the components below the bridge deck are considered ineffective, regardless of whether they have sustained any damage or not. Deck drains
through curb openings, where the water from the decks typically drains onto superstructure
elements, are an example of poor deck drainage. Decks with ponding issues may need to be
individually scored.
Ponding or ineffective drainage

10 points

No problems noted

0 points

D.5 Use of Open Decking


Reason(s) for Attribute. The presence of an open deck increases the likelihood that corrosion of the steel superstructure will occur. An open deck allows water, de-icing chemicals, and
other debris to fall directly onto the superstructure instead of running into deck drains and then
to downspout pipes, as they would in a closed deck system. As a result, the likelihood of damage
occurring in superstructure elements, bearing, and substructure elements is greatly increased.
Users may also use this as a screening attribute.
Assessment Procedure. The attribute is scored based on whether or not the bridge contains
an open deck. Common types of open decks include timber or open grating decks.
Bridge has an open deck

20 points

Bridge does not have an open deck

0 points

D.6 Year of Construction


Reason(s) for Attribute. This attribute reflects the influence of bridge age and historic
design on the most prevalent aging mechanisms in highway bridgesdeterioration of concrete
associated with corrosion of embedded reinforcement, and corrosion damage and/or fatigue
and fracture for steel structures.
The corrosion of embedded reinforcing steel occurs due to the penetration of chlorides, water,
and oxygen to the level of the reinforcement. For intact concrete, the penetration of the chlorides

70 Proposed Guideline for Reliability-Based Bridge Inspection Practices

is presently modeled as a diffusion process, using Ficks Law, which depends on time, temperature, the permeability of the concrete, and the concentration of chlorides at the components
surface. Additionally, if the concrete has suffered damage, such as cracking or spalling, chlorides
can more easily concentrate at the reinforcement, effectively expediting the corrosion process.
The quality of the concrete used in bridge construction has generally improved over time
due to concrete technology innovation, improvements in quality control, and in better supplier understanding of optimal material selection for strength and durability. Therefore, it is
reasonable to expect that a concrete component constructed to modern standards is likely to
have improved corrosion resistance characteristics compared to older components. Additionally, older structures have been exposed to the surrounding environment for a longer period of
time, and are therefore more likely to be affected by corrosion.
With respect to steel girders, the year the bridge was designed can provide valuable information
about the susceptibility of the bridge to fatigue cracking and fracture. Over the years, there have
been numerous changes in design specifications that have resulted in the improved fatigue and
fracture resistance of bridges. Four key dates have been identified; 1975, 1985, 1994, and 2009,
with regard to changes in design specifications. These dates were selected for the following reasons:
1975
Fatigue
The modern fatigue design provisions, based on the research of Fisher and others, were fully
incorporated into the AASHTO Specifications with the 1974 Interims. The basic detail categories
have not changed significantly since their introduction. Hence, 1975 was selected as a differentiator regarding fatigue design of steel bridges. Prior to 1975, fatigue design was based on principles that were not generally appropriate for welded structures. Although these early provisions
appeared in the 1965 version of the specifications and were in place through 1976, it was felt that
it was reasonably conservative to ignore the earlier provisions and set the cutoff date at 1975.
Fracture
In 1974, partly in response to the Point Pleasant Bridge collapse (1967), mandatory Charpy
V-Notch (CVN) requirements were set in place for welds and base metals as a part of the
AASHTO/AWS Fracture Control Plan. The purpose of these CVN requirements was to ensure
adequate fracture toughness of materials used in bridges. Furthermore, modern fatigue design
provisions, based on the research of Fisher and others, were fully incorporated into the AASHTO
Specifications as previously discussed. Hence, 1975 was selected as a differentiator regarding
fatigue and fracture design of steel bridges.
1985
In 1985, AASHTO introduced changes to address and to prevent distortion-induced fatigue
cracking. A common example of distortion-induced fatigue cracking is web-gap cracking.
Hence, considering the specifications introduced in 1975 and 1985, bridges designed after 1985
are less likely to be susceptible to fatigue due to primary or secondary stress ranges than bridges
built prior to these revisions.
1994
In 1994, the AASHTO design specifications changed from load factor design (LFD) to load
and resistance factor design (LRFD). The LRFD method is intended to ensure greater reliability
in bridge design. There were several changes regarding the load models and the load distribution
factors used for the fatigue limit state. These changes were intended to result in a more realistic
and reliable fatigue design. Hence, for the fatigue limit state, bridges designed after 1994 would
be expected to have improved reliability.

Attribute Index and Commentary 71

2009
In 2008, language was introduced into the AASHTO LRFD Bridge Design Specifications
which directly addressed the issue of CIF. The article provided prescriptive guidance to ensure
that details susceptible to CIF are avoided. It is included in the 2009 and later versions of the
AASHTO LRFD Bridge Design Specifications.
Assessment Procedure. The year of construction is intended to characterize the years of
environmental exposure a component has experienced or the fatigue susceptibility of the design.
The suggested values are intended to put elements into four broad classes that range from very
old to relatively new. For elements that have been replaced, the year of the replacement should
be used. Elements that have been rehabilitated should use the original construction date. These
ranges are advisory; users may consider modifying these categories based on experience with
their bridge inventory or significant changes to construction practices that may have occurred
within their state. For steel-girder categories, users should consider if the design specification used in the design of the bridge matched the contemporary specifications at the time, as
described above. If, for example, the LRFD provisions of 1994 were not implemented in the state
until 2000, then the ranges should be adjusted accordingly.
Concrete Bridge Decks, Prestressed Girders, Substructures
Built before 1950

10 points

Built between 1950 and 1970

6 points

Built between 1970 and 1990

3 points

Bridge is less than 20 years old

0 points

Steel Girders, Fatigue


Bridge designed before 1975/unknown

20 points

Bridge designed between 1976 and 1984

10 points

Bridge designed between 1985 and 1993

5 points

Bridge designed after 1994

0 points

Steel Girders, Fracture


Bridge designed before 1975/unknown

20 points

Bridge designed between 1975 and 1984

10 points

Bridge designed between 1985 and 1993

5 points

Bridge designed between 1994 and 2008

3 points

Bridge designed after 2009

0 points

D.7 Application of Protective Systems


Reason(s) for Attribute. Protective systems such as membranes, overlays, or sealers may be
applied to the surface of a concrete element to reduce the ingress of water, which may contain
dissolved chlorides or other corrosive substances. When these corrosive materials diffuse to the
level of the reinforcement, the likelihood of reinforcement corrosion increases, which may lead
to the propagation of damage. Protective systems delay or prevent this process from occurring
thereby reducing the likelihood for future corrosion damage. Some overlays have also been
shown to delay the development of spalling as a result of an increased resistance to cracking and
an increased ability to confine delamination damage (4).

72 Proposed Guideline for Reliability-Based Bridge Inspection Practices

An overlay is defined herein as an additional layer of protective material, which is applied


on top of the concrete deck and that also serves as the riding surface. Overlays may consist of
asphalt, latex-modified concrete, low-slump dense concrete, silica fume concrete, polymer concrete, or other materials.
A membrane is defined herein as a barrier that is placed on top of the concrete deck and is
then covered by another material, which serves as the riding surface. Common membranes may
consist of hot-rubberized asphalt, resin, bitumen-based liquid, or prefabricated sheets.
Sealers are somewhat different from overlays and membranes in that they are applied thinly
to concrete surfaces and penetrate the porosity of the concrete to seal it from moisture. Initially, sealers were used to counteract freeze-thaw damage and de-icing chemical-application
related scaling. With the proper use of air-entraining admixtures, the primary purpose of sealers
changed to preventing or slowing the ingress of chlorides (5). Types of sealers include silanes,
siloxanes, silicates, epoxies, resins, and linseed oil.
Surface coatings such as epoxy, polyurethane, or polyurea may also be applied to the concrete
elements of a bridge in order to increase their resistance to water intrusion and consequently
reduce their probability of developing corrosion damage. The application of these coatings can
improve the durability and corrosion resistance of concrete elements.
Each of these protective systems is intended to delay or prevent corrosion damage in concrete
bridge elements. If the protective systems are effective, then the likelihood of corrosion-related
damage will be reduced compared to unprotected elements of similar design characteristics and
environmental conditions. As a result, the application of protective systems may be considered
in the reliability assessment.
Assessment Procedure. If protective systems such as membranes, overlays, or sealers have
been applied to a concrete element, their effectiveness should be evaluated based on engineering
judgment and local experience or test data along with any documented research and field testing data that is available. Important factors to consider include the effectiveness of the applied
system as well as how often that system is applied or maintained. This attribute assumes that
overlays and sealers generally have similar effects in terms of corrosion protection for the deck.
Based on their experience, users may wish to separate certain overlays or membrane systems.
For example, an owner may have experience that indicates that low-slump overlays are having a
significant effect on extending the service life of bridge decks. In that case, the owner may wish
to increase the importance of this attribute to a moderate or high level, and distribute the scoring appropriately. The suggested scoring assumes the protective system has a low importance
relative to other design characteristics.
Never applied, poor functioning, or non-functioning

10 points

Yes, penetrating sealer, crack sealer, limited effectiveness

5 points

Yes, periodically applied, effective

0 points

D.8 Concrete Mix Design


Reason(s) for Attribute. Concrete mix designs, such as those considered to be HPC,
typically have a lower permeability and a higher durability than other traditional concrete
mixes. Therefore, high performance mixes provide an increased resistance to de-icer or marine
environment-based chloride ion penetration. This in turn can increase the time to corrosion
initiation in reinforcing steel. This design attribute is intended to consider the increased durability provided by HPC mixes.

Attribute Index and Commentary 73

The permeability of a concrete mix depends on several factors including the water to cementitious ratio, the use of densifying additives, and the use of mix-improving additives. Supplementary cementitious materials such as fly ash, ground-granulated blast furnace slag, and silica fume
have been shown to reduce permeability. Additionally, a properly designed and placed concrete
mix with a lower water to cementitious ratio will have a lower permeability.
Materials and criteria that have been identified as being beneficial in enhancing the performance of concrete bridge decks can be found NCHRP Synthesis 333: Concrete Bridge Deck
Performance (5).
Assessment Procedure. The evaluation of a bridges concrete mix design should be based
on information contained in the bridges design plans and on engineering judgment. Many
different types of concrete mixtures can be considered to be high performance, therefore, users
should consider the corrosion resistance characteristics of the particular mixture and assess if
the concrete mix used is expected to provide an increased durability relative to a typical concrete
mix design. Past experience with concrete mixes of similar characteristics should be considered.
The concrete used is not considered to be high performance

15 points

The concrete used satisfies high performance conditions

0 points

D.9 Deck Form Type


Reason(s) for Attribute. Concrete decks constructed with stay-in-place (SIP) forms have
the surface of the deck soffit hidden from visual inspection. Signs of corrosion damage such as
efflorescence, rust staining, and cracking in the deck soffit cannot typically be observed. As a
result, there can be increased uncertainty in the condition of the deck determined through visual
inspection. This attribute is intended to consider the increased level of uncertainly in the deck
condition that may exist when SIP forms are used.
Assessment Procedure. This attribute is assessed based on whether the deck has SIP forms.
SIP forms

10 points

Removable forms

0 points

D.10 Deck Overlays


Reason(s) for Attribute. Similar to SIP forms, deck overlays prevent the visual observation
of the deck condition. Signs of deterioration, corrosion damage, and cracking of the deck cannot
typically be observed. As a result, there can be increased uncertainty in the condition of the deck
determined through visual inspection. This attribute is intended to consider the increased level
of uncertainty in the deck condition that may exist for decks with overlays.
Assessment Procedure. This attribute is assessed based on whether or not the deck has an
overlay.
Deck has an overlay

10 points

Bare deck

0 points

D.11 Minimum Concrete Cover


Reason(s) for Attribute. This attribute is intended to consider the improved corrosion resistance and the increased durability associated with adequate concrete cover, and the

74 Proposed Guideline for Reliability-Based Bridge Inspection Practices

historically poor performance of bridge elements with inadequate cover. The depth of concrete
cover characterizes how far corrosive agents need to travel in order to reach the embedded steel
reinforcement. Several studies have identified that the depth of concrete cover over the top reinforcing steel mat is the most significant factor contributing to the durability of decks (5). The
importance of adequate concrete cover is also an important durability factor for other concrete
elements. The value used for this attribute should be the actual amount of concrete cover, which
may not necessarily be the design cover. If quality control procedures are adequate to ensure that
the design cover matches the as-built cover, the design cover may be used. If such quality control
procedures have not been utilized or have historically been inadequate, it may be necessary to
assess the as-built cover.
In 1970, the general recommendation for concrete cover was a minimum clear concrete cover
of 2 inches over the top-most steel. Currently, the AASHTO Standard Specifications for Highway
Bridges (2002) requires a minimum concrete cover of 2.5 inches for decks that have no positive corrosion protection and are frequently exposed to de-icing chemicals. Positive corrosion
protection may include epoxy coated bars, concrete overlays, and impervious membranes. The
AASHTO LRFD Bridge Design Specifications (2004) also requires a minimum concrete cover of
2.5 inches for concrete that is exposed to de-icing chemicals or on deck surfaces that are subject
to stud or chain wear. The concrete cover may be decreased to 1.5 inches when epoxy coated
reinforcement is used.
It is also important to note that the type of damage and the rate of damage development vary
with the amount of concrete cover. It has been reported that the type of damage changes from
cracks and small, localized surface spalls to larger delaminations and spalling as the concrete
cover increases (4). There is also an increase in the time to corrosion initiation and a reduction
in the rate of damage development when cover increases, as shown schematically in Figure E1.
In summary, as concrete cover increases, the time to corrosion initiation increases due to the
increased depth that chloride ions must penetrate to initiate the corrosion process. As corrosion
progresses, an increased concrete cover provides confinement that reduces the rate and the type
of damage that develops at the surface of the concrete element.
It should be noted that concrete cover greater than 3 inches can result in increased cracking,
providing pathways for the intrusion of water and chlorides. This may be a consideration in
special cases in which the concrete cover is unusually large.

Figure E1. Effect of concrete cover on the


time to corrosion initiation and development
of damage (4).

Attribute Index and Commentary 75

Assessment Procedure. This attribute is scored based on the actual, physical clear cover
which with the specified bridge element operates. The user should consider whether quality
control practices used at the time of construction were adequate to provide confidence that the
as-built concrete cover conforms to the design concrete cover, or if there are indications that the
concrete cover may not be adequate. In these cases, the as-built concrete cover may be required
and can be easily obtained using a covermeter.
1.5 inches or less, unknown

20 points

Between 1.5 inches and 2.5 inches

10 points

Greater than or equal 2.5 inches

0 points

D.12 Reinforcement Type


Reason(s) for Attribute. This attribute is intended to characterize whether or not the
embedded reinforcing steel has a barrier to protect it against corrosion. The most commonly
used barrier is an epoxy coating; however, galvanized bars and stainless steel, either as cladding
or as solid bars, have also been used.
Uncoated steel reinforcement will corrode easily and significantly when under attack from
corrosive elements such as chloride ions, oxygen, and water. Since this exposure is inevitable
in an operating structure, one way to slow the corrosion process is to coat the mild steel bars
with either an organic or a metallic coating or to use an alternate solid metal bar, such as stainless steel. These coatings or alternate bars help slow the corrosion process by providing either a
physical or a metallurgical barrier against the action of the corrosive elements.
The most commonly used barrier coating is fusion-bonded epoxy powder. This type of coating has been used since 1973 and has been the subject of a significant body of research. It has
been shown that, in reinforced concrete decks, if only the top mat is coated, for every year
required to consume a given amount of mild steel, it will take 12 years for the epoxy coated bar
to lose that same amount of metal. If both the top and bottom mats are coated, it may take up
to 46 years (6). This significant increase when both mats are coated is due to increased electrical
resistance, which further slows corrosion.
Two of the more common metallic coatings used are zinc and stainless steel. Zinc coated bars
are also known as galvanized bars. Conflicting reports have been given on the performance of
galvanized bars, mostly with respect to varying levels of the water to cement ratio and to whether
or not galvanized bars are used in conjunction with mild steel bars. Research suggests that galvanized bars may add 5 more years to the 10 to 15 years required for corrosion-induced stress to
manifest in unprotected bridge decks (6).
Solid stainless steel or stainless steel clad mild steel bars have also been used, although to a lesser
extent due to their higher costs. Research conducted by the State of Virginia compared the performance of stainless steel clad and stainless steel bars with uncoated carbon steel bars. The research
concluded that defect-free stainless steel clad bars performed nearly identically to the solid stainless
steel bars. These types of bars were determined to tolerate at least 15 times more chloride than the
carbon steel bars (6).
Regardless of the specific coating or reinforcement material used, protected bars generally
have a higher resistance to corrosion damage than uncoated, mild steel bars. As such, the scoring
for this attribute considers only if the rebar is protected by one of these methods, or if it is not.
Assessment Procedure. The type of reinforcement is scored based on the presence of barrier coatings or the use of alternative metal for the embedded reinforcement. This information

76 Proposed Guideline for Reliability-Based Bridge Inspection Practices

can typically be identified from the structures design plans. If suitable information is unavailable, engineering judgment should be used.
Reinforcement is uncoated carbon steel

15 points

Reinforcement has a protective coating or is


produced from an alternate corrosion
resistant metal (e.g., stainless steel)

0 points

D.13 Built-Up Member


Reason(s) for Attribute. Many bridges, especially older structures, contain built-up
members. These built-up members are sometimes more susceptible to corrosion than normal rolled steel sections because they contain pockets or crevices, which can retain water,
salt, debris, etc. This has been known to result in an accelerated corrosion rate since debris
and moisture can remain trapped. Bridge washing, if thoroughly performed, can mitigate
these effects.
Assessment Procedure. For this attribute, a built-up member refers to riveted or bolted
members. Welded members should not be included in this assessment because they do not contain the type of pockets or crevices that can trap corrosion inducing materials.
Element is a built-up member

15 points

Element is not a built-up member

0 points

D.14 Constructed of High Performance Steel


Reason(s) for Attribute. In addition to possessing higher yield strengths than normal steels,
high performance steels (HPSs) generally have greater fracture toughness than that required by
ASTM A709, and of other common bridge steels. Improved fracture toughness results in steel
that is more resistant to fracture than normal steels. This is because it is more likely that cracks
will propagate at a slower rate, and could even arrest, in HPS compared to normal steels.
At this time, the CVN levels required for HPS in ASTM A709 are not established with the
objective of achieving any particular level of fracture resistance or crack tolerance. Hence, the
benefits provided by using HPS, if the steel just meets the ASTM A709 specification, are limited.
Therefore, the suggested ranking of HPS is low in terms of contribution to durability and reliability (10 pts), relative to normal steel. This may change as future research becomes available
and the minimum required CVN values increase for HPS.
Assessment Procedure. This attribute should be scored based on whether or not the element is constructed out of HPS. If there is no documentation or it is unknown if the element is
constructed of HPS, the attribute should be scored accordingly.
Element is not constructed of HPS/unknown

10 points

Element is constructed of HPS

0 points

D.15 Constructed of Weathering Steel


Reason(s) for Attribute. Weathering steel is a type of steel that contains alloying elements
that increase the inherent corrosion resistance of the steel. For this reason, weathering steels are
less susceptible to corrosion than normal black steels. However, this is only true if the steel is
used in the proper environment and is detailed properly.

Attribute Index and Commentary 77

Assessment Procedure. This attribute is scored based on whether or not the element is
constructed using weathering steel and is detailed and located in a manner that minimizes the
contact of the steel with de-icing chemicals and moisture. If it is unknown if the element is composed of weathering steel, the element should be scored accordingly. The assessment procedure
assumes that the steel is used in the proper environment and is detailed properly. Guidance
on the appropriate application of uncoated weathering steel can be found in FHWA Technical Advisory T-5140.22 (7). The document also includes recommendations for maintenance to
ensure continued successful performance of the steel.
Element is not constructed of weathering
steel or location and detailing may allow
impact of ambient or de-icing chemicals
on steel surfaces

10 points

Element is constructed of weathering steel


and properly detailed consistent with
FHWA Technical Advisory T-5140.22

0 points

D.16 Element Connection Type


Reason(s) for Attribute. Welded connections are usually more susceptible to the effects of
fatigue damage than other types of connections, as there is a direct path for cracks to propagate
between connected elements. For example, a crack in a flange can grow into the web through
the web-to-flange weld. Fatigue cracking is generally of greatest concern for welded details that
have low fatigue resistance, such as D, E, and E, along with residual stresses and weld toe defects.
Riveted connections, unlike welded connections, do not offer a direct path for cracks to propagate from one element to another. Using the web-to-flange connection example, cracks in an
angle used to make up a flange are not able to grow directly into the web plate because the elements are not fused together. Hence, there is a certain amount of redundancy at the member
level. Nevertheless, the quality of the rivet hole (e.g., punched vs. drilled) and a lack of consistent
pretension in rivets results in these details being classified as category D.
Similar to riveted connections, high strength (HS) bolted connections are more resistant
to a fatigue crack propagating from one component of a member to another, as compared to
welded members. A properly tightened HS bolt generates very high compressive forces in the
connection. The pretension force is much greater and is much more consistently achieved in
a HS bolted connection than in a riveted connection. As a result of the significant pretention
in a fully tightened A325 or A490 bolt, the quality of the hole itself has little or no effect on the
fatigue resistance of the connection (in contrast to riveted joints). As a result, they are classified
as category B details.
It is noted that considering the element connection type may appear to be a double penalty
when considered in conjunction with D.17 Worst Fatigue Detail Category. However, it is clear
that should cracking occur at a welded detail in a main member, it is more likely to become an
issue than in, say, the equivalent bolted detail simply due to the fact that there is no direct path
for cracks to grow from component to component in the bolted joint. Hence, it is considered
a better condition even though both welded and bolted details may both be classified as
categoryB. Riveted details, which do not have as high a fatigue resistance as HS bolted connections, but are not as susceptible to crack propagation as welded joints, have been arbitrarily
scored in the middle.
Assessment Procedure. If the element has multiple types of connections, the worst type of
connection should be scored for this attribute.

78 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Element connected with welds

15 points

Element connected with rivets

7 points

Element connected with HS bolts

0 points

D.17 Worst Fatigue Detail Category


Reason(s) for Attribute. The likelihood of fatigue cracking is influenced by the type of
fatigue detail category present. It is generally accepted that poor fatigue details are more likely
to develop cracks than more fatigue resistance details. This is implied in the current AASHTO
LRFD Bridge Design Specifications, which discourages the use of details lower than category C and
encourages design for infinite life. Fortunately, since the introduction of the modern AASHTO
fatigue provisions in 1975, the use of poor details (D, E, and E) has been greatly reduced. Hence,
details in bridges designed over the past 30 years or so will typically be of higher fatigue resistance.
Assessment Procedure. The worst type of detail subjected to tensile stress ranges in the
element or member should be used for this attribute. The AASHTO fatigue details A through
E should be used.
Fatigue detail category E or E

20 points

Fatigue detail category D

15 points

Fatigue detail category C

5 points

Fatigue detail category A, B, or B

0 points

If the element has multiple types of connections, the worst type of connection should be
scored for this attribute.
Element connected with welds

15 points

Element connected with rivets

7 points

Element connected with HS bolts

0 points

D.18Skew
Reason(s) for Attribute. Bridge skew can introduce unanticipated forces in a bridge deck,
deck joints, and superstructures. Thermal expansion of the superstructure and deck may introduce uneven strain distributions and/or torsional forces. As a result, bridges with high skew
angles may suffer atypical deterioration patterns including cracking in bridge decks, failure of
joints and bearing, and distortion-induced cracking at diaphragms (812).
Assessment Procedure. This attribute is typically scored based on the recorded skew angles
for a bridge. Angles of 30 degrees or greater may be used as a value for evaluating the potential
for adverse skew angle effects. This attribute may also be used as a screening attribute.
Skew 30 or more

20 points

Skew 2030

10 points

Skew less than 20

0 points

D.19 Presence of Cold Joints


Reason(s) for Attribute. Cold joints or construction joints within deck spans can sometimes result in leakage of water and de-icing chemicals through the deck and onto the supporting

Attribute Index and Commentary 79

superstructure. This may result in accelerated deterioration patterns including coating failure
and section loss for steel members, corrosion damage in concrete members, and/or corrosion
damage in the deck.
Assessment Procedure. This attribute is typically scored based on the presence of known
cold joints within the deck span. Data to support this assessment may come from inspection
reports, because cold joints that are performing as designed may not be known.
Presence of cold joints

10 points

No known cold joints

0 points

D.20 Construction Techniques and Specifications


Reason(s) for Attribute. Construction techniques and specifications have evolved over
time to improve the durability and performance characteristics of bridges. Certain construction techniques and specifications used during previous eras may be problematic, and result in
deterioration and damage patterns that can be associated with the techniques or specification
in use at the time of bridge construction. For example, reduced bridge deck thickness may have
been typical during a certain era. Over time, the reduced deck thickness may be shown to reduce
the durability of the bridge deck and result in deck damage such as punch-through. As a result,
decks constructed during that era may be more likely affected by a certain damage mode than
bridges constructed during other eras.
Assessment Procedure. This attribute will typically be identified by RAP members based
on experience of bridge inspection and maintenance personnel. Historical records documenting
the evolution of design standards and construction techniques may be necessary to identify the
specific era, or estimates based on experience may be used. This attribute may also be used as a
screening attribute.
Bridge constructed during identified era

20 points

Bridge not constructed during identified era

0 points

D.21 Footing Type


Reason(s) for Attribute. Spread-type footings may be susceptible to the adverse effects of
scour, soil sliding, or rotations due to uneven settlement or subsidence. In contrast, pile foundations may be unaffected by these phenomena. As such, deterioration patterns and damage
modes that affect spread footings may not be relevant for pile foundations.
Assessment Procedure. This attribute can typically be determined from the design drawing
available in the bridge file. This attribute may be used as screening criteria for specific damage
modes that affect spread footings, but would not affect pile foundations.
Spread-type footing

15 points

Pile foundation

0 points

D.22 Subsurface Soil Condition


Reason(s) for Attribute. Footings on certain soils may be susceptible to the effects of soil
sliding or rotations due to uneven settlement or subsidence. This attribute is typically utilized in
conjunction with D.21 to reflect the increased likelihood of damage modes such as substructure
rotations, cracking, or displacements for bridges in certain geographic regions.

80 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Assessment Procedure. Subsurface soil conditions susceptible to these effects are typically
known to geotechnical engineers and/or maintenance personnel. This attribute may be identified based on soil testing results or experience.
Poor or unknown subsurface soil conditions

20 points

Acceptable soil condition or pile foundations

0 points

Loading Attributes
L.1ADTT
Reason(s) for Attribute. The ADTT on a bridge is used to characterize the frequency of
occurrence of large external loads on the bridge due to heavy vehicles. Large transport trucks
or other heavy vehicles place stress on a bridge as static and dynamic loads, the latter reflecting
impact and other dynamic amplification effects.
As ADTT levels increase, the rate of damage formation and accumulation in concrete is typically expected to increase. This is in part because the stresses caused by traffic loads accelerate the
effects of the internal expansion forces from reinforcement corrosion (4). These loads, especially
when placed on a bridge with existing deterioration, will open cracks and possibly allow corrosive elements to enter the cracks or increase the crack density. Experience has shown that bridge
decks exposed to heavy truck traffic generally deteriorate at a much higher rate than decks with
little or no truck traffic.
For steel girders, research has shown that trucks produce nearly all of the fatigue damage in
highway bridges. Hence, a bridge with high truck traffic (high ADTT) will have a higher probability of fatigue damage. Of course, the converse is also true, bridges with little or no truck
traffic (e.g., HOV bridges) are unlikely to experience fatigue cracking.
It is important to note that ADTT only considers the load side of the equation. The likelihood of fatigue cracking also depends on the resistance side of the equation, which is addressed
by the D.16 Element Connection Type and D.17 Worst Fatigue Detail Categories. Although
ADTT does not provide an exact correlation to the stress ranges an element will experience, it
does provide a reasonably good understanding of how quickly fatigue damage may accumulate.
Assessment Procedure. This attribute should be scored based on the ADTT.
For steel structures, the scoring limits for ADTT were taken from a recent study on fracture
critical bridges titled A Method for Determining the Interval for Hands-On Inspection of Steel
Bridges with Fracture Critical Members (13). Although these limits were developed primarily with
fracture critical bridges in mind, it was decided these limits could be applied to other highway
bridges as well for the fatigue limit state. The reasoning behind the limits as documented in Parr
and Connors report is as follows:
The ADTT limit of 15 comes from the fact that for bridges where the ADT is less than 100,
the ADT is generally not reported in the NBIS. During the Purdue University Workshop, it was
agreed than an ADTT of 15% (of the ADT) was a reasonably conservative estimate of the proportion of trucks crossing a typical low volume bridge. Hence, 15% of the lowest ADT reported
in the NBIS (ADT = 100) yields an ADTT of 15.
The lower bound value of 100 was set such to separate bridges in rural areas versus moderately traveled bridges. The upper bound limit of an ADTT equal to 1,000 was obtained by simply
increasing the moderate limit by a factor of 10. It was included simply to create a boundary
between heavily and moderately traveled bridges.

Attribute Index and Commentary 81

For concrete bridges, high ADTT will likely have the most significant effect on the durability of the bridge deck. Superstructure components will be affected to a much lesser extent; if
designed to modern standards, high ADTT may have little effect on the durability of superstructure components. Deck joints may also deteriorate more rapidly in the presence of high ADTT.
Users may wish to adopt different thresholds for the scoring model, depending on typical
traffic patterns and needs.
Concrete Bridge Deck, Prestressed Concrete Girder
ADTT is greater than 5,000

20 points

ADTT is moderate

10 points

ADTT is minor

5 points

No heavy trucks

0 points

Steel Girders
ADTT is greater than 1,000

20 points

ADTT is between 100 and 1,000

15 points

ADTT is between 15 and 100

5 points

ADTT is less than 15

0 points

L.2 Dynamic Loading from Riding Surface


Reason(s) for Attribute. This attribute is intended to consider the detrimental effects of
dynamic loading on the deterioration patterns for concrete bridge decks. This attribute would
typically be used to adjust assessments to consider a reduction of the durability of bridge decks
with high dynamic loads (i.e., high speed traffic and high ADTT). This attribute is included to
consider cases where the riding surface or the deck joint becomes damaged, such as through
the development of potholes, rough patches, or a bump at the end of the bridge, and increased
dynamic forces are created due to the traffic loading. These forces place additional stress on the
structure leading to a perpetual cycle of damage propagation that accelerates the rate of deterioration for the deck element (14).
Assessment Procedure. This attribute is based on engineering judgment. Considerations
in assessing this attribute include the roughness of the riding surface, the existence of potholes
and patches, durability of deck joints, ADTT, and traffic speeds.
Dynamic forces leading to increased rate of
deterioration a significant consideration

15 points

Dynamic forces not a significant consideration

0 points

L.3 Exposure Environment


Reason(s) for Attribute. The environment surrounding a bridge can have a significant
effect on the rate of deterioration, particularly for corrosion. This attribute is intended to characterize the macro-environment surrounding a bridge and account for the likelihood of increased
deterioration rates in environments that are particularly aggressive, such as coastal or marine
environments. Aggressive environments typically have high ambient levels of chlorides, high
ambient moisture levels (high humidity or frequent wet/dry cycles, increased temperature),
and the presence of other harmful chemicals (i.e., high levels of carbon dioxide, sulphates, etc.).

82 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Assessment Procedure. The assessment procedure is similar to other environmental exposure classifications that are already in practice. Marine environments are deemed to be the most
severe due to the high levels of ambient chlorides and moisture. Moderate environments are
those in which corrosive agent levels (water and chlorides) are elevated but lower than those
found in marine or other severe exposures. Industrial environments are less severe than
marine but may contain other harmful chemicals. Under modern regulatory constraints, airborne pollutant levels associated with industrial environments are minimized, and this should
be considered in the assessment of industrialized environments. Benign environments are
those in which application of de-icing chemicals is minimal or nonexistent; the environments
may be arid and atmospheric pollutants typical.
Severe/Marine

20 points

Moderate/Industrial

10 points

Benign

0 points

L.4 Likelihood of Overload


Reason(s) for Attribute. This attribute can be used when the likelihood of overload is a
consideration for the bridge or a family of bridges being assessed. The likelihood of overload
is used to characterize the chance that a bridge will be loaded beyond its inventory load rating.
Such overloads generally increase the deterioration rate for structural elements. The probability
of this occurring may be greater for bridges with a reduced capacity, such as those that have
already been load posted.
Assessment Procedure. This attribute is scored based on how likely it is that a bridge will be
overloaded. Sound engineering judgment should be used to assess this attribute.
High likelihood of overload

15 points

Moderate likelihood of overload

10 points

Low likelihood of overload

0 points

L.5 Rate of De-icing Chemical Application


Reason(s) for Attribute. This attribute is intended to characterize the volumes of the
de-icing chemicals containing chloride ions that are being applied regularly to the surface of
the deck. The detrimental effects of de-icing chemicals on the durability of bridge elements
are well known. The intrusion of chloride ions to the level of the reinforcing steel provides
an important driving force for corrosion of the reinforcing steel (15). When combined with
oxygen and water, higher levels of de-icing chemical application generally lead to more rapid
and severe reinforcement corrosion rates. The presence of increased chloride concentrations
at the surface of the concrete increases chloride diffusion rates, shortening the time for the
initiation of corrosion in the steel. If faulty deck joints or a substandard drainage system are
present, which permit water seepage, bridge elements below the deck may also be affected by
increased chloride ion levels. This will lead to increased levels of corrosion and consequently
to corrosion-related damage.
Assessment Procedure. This attribute can be scored based on the average annual number
of applications of de-icing chemicals to the deck surface. The application rates may either be
expressed quantitatively, if the bridge owner keeps such records, or on a qualitative scale. Factors
that could be used to help estimate the rate of salt application include the ADT of the roadway
and the amount of snowfall the bridge experiences. Typically, bridges with high ADT lie along

Attribute Index and Commentary 83

critical roadways that may receive the focus of local maintenance crews for the application of deicing chemicals. Obviously, the more frequent the snowfall, the more often de-icing chemicals
are likely to be applied. Users may have other data or information regarding the application of
de-icing chemicals that can be used to develop rationales identifying those bridges exposed to
high levels of de-icing chemicals and those where de-icing chemical use is minimal.
High (more than 100 applications per year)

20 points

Moderate

15 points

Low (less than 15 applications per year)

10 points

None

0 points

L.6 Subjected to Overspray


Reason(s) for Attribute. Overspray refers to the de-icing chemicals on a roadway that are
being picked up and dispersed by traveling vehicles onto adjacent highway structures, including
bridges and their substructures. Bridges that are located over roadways may receive overspray
from the road below. Since overspray typically consists of salt or other de-icing chemicals, more
exposure increases the likelihood of developing a corrosion problem.
It is noted that L.6 Subjected to Overspray is explicitly considered to be a separate item from
L.5 Rate of De-icing Chemical Application. This because some bridges may not have de-icing
chemicals directly applied to their decks, but still can be exposed to overspray from below. An
example of this would be a rural road over an interstate. However, to address the more severe
condition where de-icing chemicals are applied to the bridge directly and by overspray, the items
are considered separately.
Assessment Procedure. Similar to the rate of de-icing chemical application, a quantitative
estimate of overspray exposure may be difficult. The frequency of de-icing chemical application
on the highway that the bridge crosses (if applicable) can be used to aid in estimating the overspray exposure. The vertical clearance of the bridge is also a consideration. For example, a bridge
with greater than 20 feet of vertical clearance over the roadway below may experience minimal
effects from overspray. In any case, sound engineering judgment should be used. The suggested
scoring scheme is based on the generally more significant effect of overspray on steel bridge
elements. These suggested scales should be modified appropriately based on local experience.
Concrete Bridge Deck, Prestressed Girder, Substructure
Severe overspray exposure

15 points

Moderate overspray exposure

7 points

Low exposure overspray or not over a roadway

0 points

Steel Girder
Severe overspray exposure

20 points

Moderate overspray exposure

10 points

Low exposure overspray or not over a roadway

0 points

L.7 Remaining Fatigue Life


Reason(s) for Attribute. The remaining fatigue life of an element is somewhat related to the
probability of a fatigue crack propagating to the point of brittle fracture. Obviously, for elements

84 Proposed Guideline for Reliability-Based Bridge Inspection Practices

that have longer remaining fatigue lives, there is a lower probability of failure due to fatigue
cracking than for elements with shorter remaining fatigue lives.
Assessment Procedure. The remaining fatigue life of an element can be determined using
any established method. Insufficient fatigue life refers to a fatigue life that is less than the required
service life or some other interval defined by the owner (e.g., less than 10 years). It is noted that
it is possible to calculate a life of less than the length of time the bridge has been in service (i.e., a
negative fatigue life). In many cases, although a negative fatigue life has been calculated, there is
no evidence of fatigue cracking on the structure. Although a negative fatigue life does not make
physical sense, it does suggest that the probability of failure due to fatigue cracking is greater. In
such cases, more in-depth evaluation efforts are justified, such as field testing or monitoring to
obtain in-service stress range histograms or a more accurate finite element model of the structure. Often, the more in-depth evaluations reveal that there is significant remaining fatigue life.
Sufficient fatigue life refers to a fatigue life that exceeds the expected service life, or a defined
life required by the owner (e.g., 10 years until replacement) of the element, but is not infinite.
Infinite life is the case in which fatigue cracking is not expected to propagate during the life of
the structure. It is noted that a greater penalty is placed on not having any knowledge of the
remaining fatigue life than on having performed a fatigue analysis that determined a negative
fatigue life.
Unknown remaining fatigue life

10 points

Insufficient remaining fatigue life

7 points

Sufficient remaining fatigue life

3 points

Infinite remaining fatigue life

0 points

L.8 Overtopping/High Water


Reason(s) for Attribute. Certain bridges are susceptible to periodic overtopping or high
water condition in which the bridge superstructure is partially or totally immersed in water.
Such condition may not adversely affect the loading carrying capacity of the structure; however,
this condition may increase the likelihood that A) the structure is impacted by debris or ice
in the water, or B) debris is deposited on the flanges and surrounding the bearing areas of the
bridge. Impact from debris or ice in the water may increase the likelihood that a certain bridge
suffers impact damage, even though the structure is not over a roadway. Debris deposited on
the superstructure or at the bearing will retain moisture and may accelerate corrosion damage.
Assessment Procedure. Bridges that are likely to be overtopped during periods of high
water are typically documented in the NBIS data submitted annually to the FHWA. Experience
may also be used to identify bridges susceptible to the adverse effects of high water. Scoring of
this attribute may be different values for conditions A and B.
Periodic overtopping/high water

20 points

No overtopping/high water

0 points

Condition Attributes
C.1 Current Condition Rating
Reason(s) for Attribute. The condition rating for a bridge component describes the existing, in-place bridge as compared with the as-built condition. The condition ratings provide

Attribute Index and Commentary 85

an overall characterization of the general condition of the entire component. It is reasonable to assume that a given element that has already shown signs of damage is more likely
to deteriorate to a serious condition than an element showing little or no signs of damage.
It is typical for a concrete component with a condition rating of 5 or less to have observable corrosion damage in the form of cracking or spalling (either as open spalls or patched
spalls). Such damage provides pathways for the increased penetration of chlorides ions and
for increased rates of damage accumulation. For steel elements, low condition ratings are
frequently emblematic of significant corrosion damage. Fatigue cracking or member distortions due to unexpected settlement, etc. may be present. Conversely, components with a high
condition rating (6 or above) typically have lower levels of existing deterioration. Consequently, some consideration should be given to the overall component rating when assessing
the durability of the bridge element.
Assessment Procedure. For this attribute, a condition rating of 5 or less is considered to
have a much higher likelihood for accelerated damage than component with higher condition ratings. A condition rating of 6 is considered to have a smaller likelihood of accelerated
damage.
Condition rating is 5 or less

20 points

Condition rating is 6

5 points

Condition rating is 7 or greater

0 points

C.2 Current Element Condition State


Reason(s) for Attribute. When element-level inspections are conducted under the AASHTO
Bridge Element Inspection Manual, element condition states (CS) that are linked to specific evidence of damage or deterioration to the subject bridge element are defined. Elements or portions of elements in CS 1 typically have very little or no evidence of deterioration. Elements or
portions of elements in CS 2 have some evidence of damage. As such, it is reasonable to assume
that if a given element is entirely in CS 1, the likelihood of severe damage occurring in the near
future is lower than an element with portions of the element in CS 2, 3, or 4. This attribute is
intended to consider the positive attributes of an element in CS 1.
Assessment Procedure. For this attribute, the current CS for a given bridge element is considered. For elements entirely in CS 1, the scoring of 0 points is suggested, for elements where CS 3
is indicated for any portion of the element, a score of 20 points is suggested. Users may wish to
utilize appropriate gradations for elements with conditions indicated as CS 2. The severity and
the significance of CS 2 vary by element, and the RAP may wish to develop alternative scoring
schemes based on specific elements and CS apportionment. Element-level inspection implementation varies at the owner level, and therefore appropriate scoring should be considered by
the RAP according to existing inspection practices.
CS 2 is indicated for a significant portion
of the element, or CS 3 is indicated for
any portion of the element

20 points

Condition State 2 is indicated for a


minor portion of the element

10 points

Condition State 1 is indicated for


entire element

0 points

86 Proposed Guideline for Reliability-Based Bridge Inspection Practices

C.3 Evidence of Rotation or Settlement


Reason(s) for Attribute. This attribute is intended to consider the effects of unexpected
rotation or settlement of abutments and piers. Use of this attribute is for minor settlements
or rotations that do not affect the structural capacity, but may result in atypical or accelerated
deterioration patterns. Significant rotations or settlements may require engineering analysis.
The rotation of a bridge substructure beyond its design tolerances may result in damage that
is manifested by cracking, skewing, and/or misaligned bridge components. Unexpected settlements may result in cracking that provides pathways for intrusion of water and chlorides, leading to accelerated corrosion of reinforcing steel.
Assessment Procedure. Evidence of rotation or settlement should be rated based on their
severity using engineering judgment.
Rotation or settlement resulting in cracking of
concrete, misaligned joints, or misaligned members

15 points

Minor evidence of rotation or settlement with the


potential to result in unexpected cracking or poor
joint performance

5 points

No evidence of rotation

0 points

C.4 Joint Condition


Reason(s) for Attribute. The presence of one or more leaking joints will dramatically
increase the possibility for corrosion related deterioration on the elements below the deck.
This is because joints that are leaking will usually leak chloride-contaminated water directly
onto other bridge components such as the superstructure, substructure, and bearing areas. This
allows corrosion to initiate and propagate at a faster rate in the affected elements.
Assessment Procedure. This attribute should be rated based on either visual observation
or on information contained in bridge inspection reports. For this attribute, the presence of a
leaking joint is considered to be severe. If a joint has become debris filled, there is an increased
probability that that joint will become damaged and start to leak in the near future. Users should
consider historical experience with typical joints in their inventory in evaluating this attribute.
For example, if certain typical joint types are expected to have a service life of less than 5 years,
it may be appropriate to assume that this joint is a leaking joint, because even if it is not leaking
currently, it is expected to leak in near future. Open joints should be expected to allow for the
passage of water and debris, and thus should be scored accordingly if this effect is unmitigated.
For bridges that are jointless, it is assumed that the bridge is performing as intended and deck
drainage is not affecting the bearing areas.
Significant amount of leakage at joints

20 points

Joints have moderate leakage or are debris filled

15 points

Joints are present but not leaking

5 points

Bridge is jointless

0 points

C.5 Maintenance Cycle


Reason(s) for Attribute. This attribute is intended to consider the positive benefits of
consistent maintenance and preservation activities on the durability and the reliability of
bridge elements. Activities such as deck cleaning, maintenance of drainage, debris removal,

Attribute Index and Commentary 87

washing out joints, and periodic application of the sealers help preserve bridge elements
and extend their service lives. Conversely, a bridge that does not receive periodic maintenance and preservation activities is likely to experience damage and deterioration much
earlier in its service life, and deteriorate at a higher rate relative to a bridge receiving consistent,
periodic maintenance.
Assessment Procedure. This attribute is scored based on the bridge maintenance policies
and practices within the particular inventory being assessed. The RAP panel should consider
the policies and practices within its state with regard to the intensity of maintenance activities within particular regions, districts, or municipalities. For example, state-owned bridges
typically receive more consistent and thorough maintenance than locally-owned bridges.
Bridges located in rural areas may receive less intense maintenance than those located near
population centers, etc. The RAP should consider specific situations within its bridge inventory when assessing this attribute, and develop criteria for establishing which bridges receive
regular maintenance that can be expected to prevent deterioration, and those bridges which
do not.
Bridge does not receive routine maintenance

20 points

Some limited maintenance activities

10 points

Bridge is regularly maintained

0 points

C.6 Previously Impacted


Reason(s) for Attribute. If a bridge has been previously struck or impacted by a vehicle,
it is reasonable to assume that there is an increased probability of further impact damage. The
element could also have been damaged as a result of previous impact, which has been shown to
decrease, for example, a steel girders resistance to brittle fracture (16). For concrete bridge elements, impacts can compromise the concrete cover, resulting in the exposure of embedded steel
elements. The occurrence of previous impacts should be considered in the analysis for potential
impact damage.
Assessment Procedure. This attribute is scored based only on whether or not the bridge
has been previously impacted. If the impact risks have been mitigated, this should be considered
in the analysis.
Bridge has been previously impacted

20 points

Bridge has not been previously impacted

0 points

C.7 Quality of Deck Drainage System


Reason(s) for Attribute. The purpose of the deck drainage system is to get water, de-icing
chemicals, and debris off of the bridge deck effectively, without draining directly onto other elements of the bridge, such as the superstructure and the substructure elements. This attribute is
intended to address leakage or deck drainage onto other bridge elements as a result of damage,
deterioration, or the ineffective performance of a deck drainage system. Deck drainage systems with
ineffective designs would typically be address using attribute D.4 Poor Deck Drainage and Ponding.
Assessment Procedure. This attribute is based on the performance of the drainage system
in place on the bridge deck. Since estimating the quality of the drainage system is subjective, it
should be based on experience, engineering judgment, and common sense. Some key factors to

88 Proposed Guideline for Reliability-Based Bridge Inspection Practices

consider when scoring this attribute include build-up at the deck inlet grates, clogged drains or
pipes, section loss in pipes, etc.
Deck drains directly onto superstructure or substructure
components, or ponding on deck results from poor drainage

20 points

Drainage issues resulting in drainage onto superstructure


or substructure components, or moderate ponding on deck;
effects may be localized

10 points

Adequate quality

0 points

C.8 Corrosion-Induced Cracking


Reason(s) for Attribute. This attribute considers the presence of corrosion-induced cracking in concrete bridge elements. Corrosion-induced cracking typically occurs due to the expansion of reinforcing steel caused by the development of corrosion by-products on the surface of
the bar. This expansion leads to cracking of the concrete, providing pathways for water and chlorides to penetrate to the reinforcement level. Frequently, this type of cracking is accompanied
by rust staining. Such evidence of active corrosion would typically be detected during a typical
visual inspection of a bridge. The presence of active corrosion increases the likelihood for corrosion damage to occur to a severe extent in the future.
Assessment Procedure. This attribute is scored based on the presence and the severity of
corrosion-induced cracking in concrete bridge elements. The determination of the significance
of the cracking should be based on engineering judgment.
Significant corrosion-induced cracking

20 points

Moderate corrosion-induced cracking

10 points

Minor corrosion-induced cracking

5 points

No corrosion-induced cracking

0 points

C.9 General Cracking


Reason(s) for Assessment. This attribute is used to characterize the presence nonstructural cracks in concrete. These cracks may result from shrinkage, thermal forces, or other
non-structural effects. These cracks can provide pathways for the intrusion of chlorides to the
level of the reinforcement. It is generally recognized that cracks perpendicular to the reinforcing
bars hasten the corrosion of the intersected reinforcement by facilitating the ingress of moisture,
oxygen, and chloride ions. Cracks that follow the line of a reinforcing bar are much more serious,
since the length of the bar equal to the length of the crack is exposed to corrosive elements. The
presence of cracking also reduces the concretes ability to contain spalling as the reinforcement
corrodes. This attribute is generally used for cracking other than corrosion-induced cracking,
which is described in attribute C.8.
Assessment Procedure. The rating of this attribute depends on engineering judgment.
More specific guidance to classifying crack sizes and density can be found in the 2010 edition of
the AASHTO Bridge Element Inspection Manual.
Widespread or severe cracking

15 points

Moderate cracking present

10 points

Minor or no cracking present

0 points

Attribute Index and Commentary 89

C.10Delaminations
Reason(s) for Attribute. Delaminations are subsurface cracks in concrete generally parallel
to the concrete surface. Delaminations are caused by the formation of horizontal cracking as a
result of volumetric expansion of the reinforcing steel during the corrosion process. Delaminations are typically emblematic of the corrosion of embedded steel, and thus provide an early
indicator of where future spalling is likely to occur. This attribute is intended to consider that
concrete elements with delaminations are more likely to experience deterioration and damage in
the future, relative to elements in which delaminations are not present. The detection of delaminations in concrete can reduce the uncertainty in determining if there is active corrosion that is
manifesting in damage to the concrete.
This attribute may also be used to characterize conditions for a deck overlay. Under these
conditions, delaminations are indicative of a loss of bond between the overlay and the substrate.
Overlays that are debonding are likely to deteriorate more rapidly than an overlay with good
bonding characteristics.
It is implied that some form of NDE has been conducted to address this attribute, as delaminations are not visibly detectable. This typically includes hammer sounding or chain drag, but
may include other techniques such as infrared thermography, impact echo, or other methods.
Assessment Procedure. This attribute is scored based on inspection results that indicate
the extent of delaminations present in a given concrete element. This attribute should be scored
based on the amount of surface area of the structure that includes delaminations. Suggested
values for the significant levels of delamination are indicated below.
Significant amount of delaminations present
(greater than 20% by area) or unknown

20 points

Moderate amount of delaminations present


(5% to 20% by area)

10 points

Minor, localized delaminations


(less than 5% by area)

5 points

No delaminations present

0 points

C.11 Presence of Repaired Areas


Reason(s) for Attribute. Repaired spalls and patches are a way to temporarily seal reinforcement exposed as a result of damaged concrete. However, even though the reinforcement is
again sealed from the environment, the existing corrosion can continue to propagate. Patches
frequently have a relatively short service life, especially when traffic loading is high.
The service life of deck patches ranges from 4 years to 10 years (17), although an FHWA
TechBrief indicates that the service life of a patch ranges from 4 years to only 7 years (18). The
service life of the patch depends largely on the corrosivity of the surrounding concrete and the
development of the halo effect. When concrete is contaminated with chlorides in concentrations
greater than the threshold level in the area surrounding the patches, inadvertent acceleration of
the corrosion rate can occur. The patched area acts as a large non-corroding site (i.e., cathodic
area) adjacent to corroding sites (i.e., anodic areas), and thus corrosion cells are created.
Assessment Procedure. The presence of repaired areas should be scored based on the total
surface area of the bridge that has repaired areas. Engineering judgment should be exercised. If
the repaired areas result from impact damage or other non-corrosionrelated damage, and chlorides levels for the intact concrete are expected to be nominal, a reduced score may be assigned.

90 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Significant amount of repaired areas

15 points

Moderate amount of repaired areas

10 points

Minor amount of repaired areas

5 points

No repaired areas

0 points

C.12 Presence of Spalling


Reason(s) for Attribute. This attribute is intended to consider the presence of spalling on
concrete bridge elements. Open spalls are sections of concrete that have separated from the
larger mass of concrete and fallen off of the structure, usually exposing the underlying reinforcement. Unrepaired spalling allows corrosive elements to directly contact the exposed reinforcement and prestressing steel, if present. This will lead to accelerated rates of corrosion damage in
the area surrounding the spall.
Users may wish to include repaired spalls under this attribute, or utilize the attribute C.11
Presence of Repaired Areas.
Assessment Procedure. This attribute is scored based on the severity and the extent of
spalling as reported in bridge inspection reports. Users should consider the importance of the
spalling in terms of the structural performance of the element under consideration in developing their scoring methodology. Spalling that leads to the exposure of prestressing strands
is considered significantly more important than spalling in a reinforced element exposing the
mild steel bars.
Significant spalling (greater than 10% of area
with spalling, rebar or strands exposed)

20 points

Moderate spalling (greater than 1 inch deep or


6 inches in diameter or exposed reinforcement)

15 points

Minor spalling (less than 1 inch deep


or 6 inches in diameter)

5 points

No spalling present

0 points

C.13Efflorescence/Staining
Reason(s) for Attribute. This attribute is intended to consider the increased likelihood
of corrosion damage associated with the presence of efflorescence on the surface of concrete
elements. Efflorescence is a white stain on the face of a concrete component which results
from the crystallization of dissolved salts. While efflorescence is typically considered an aesthetic problem, it may be indicative of a problem with the concrete mix and may contribute
to corrosion initiation. Efflorescence on the soffit of a bridge deck typically indicates that
water is passing freely through the deck, likely carrying with it chlorides that may cause corrosion of the reinforcing steel. When rust stains are present, the corrosion of reinforcing
steel is assured.
Extensive leaching causes an increase in the porosity and the permeability of the concrete,
thus lowering the strength of the concrete and making it more vulnerable to hostile environments (e.g., water saturation and frost damage, or chloride penetration and the corrosion of
embedded steel). Those concretes that are produced using a low water-cement ratio, adequate
cement content, proper compaction, and curing are the most resistant to leaching that results in
efflorescence on the surface of the concrete (19).

Attribute Index and Commentary 91

Assessment Procedure. This attribute is scored based on inspection results. The scoring for
this attribute is based on the existence of efflorescence stains and whether or not rust stains have
also been deposited from corroding reinforcement.
Moderate to severe efflorescence with rust staining;
severe efflorescence without rust staining

20 points

Moderate efflorescence without rust staining

10 points

Minor efflorescence

5 points

No efflorescence

0 points

C.14 Flexural Cracking


Reason(s) for Attribute. When the primary load-bearing members in a concrete bridge
exhibit flexural cracking, it may indicate that the members were either inadequately designed
for the required loading, that overloads have occurred, or that deterioration has occurred that
has reduced the load-bearing capacity of the members. In any case, large flexural cracks can be
indicative of an inadequate load-bearing capacity that may require an engineering analysis in
order to determine the cause of the cracking and the resulting effect on the load capacity of the
structure. As a result, bridges exhibiting moderate to severe flexural cracking should be screened
from the general reliability assessment unless appropriate engineering analysis indicates that the
cracking is benign. Flexural cracking in a prestressed element is generally more significant than
in a reinforced concrete element.
In cases where flexural cracking is minor or appropriate assessment has indicated that the
cracking is not affecting the adequate load capacity of the element, the cracking nonetheless
may provide pathways for the ingress of moisture and chlorides that may cause corrosion of
the embedded steel. This attribute is intended to consider the increased likelihood of corrosion
resulting from the cracking in the concrete.
Assessment Procedure. Flexural cracks will typically present themselves with a vertical orientation either near the bottom flange at mid-span or near the top flange over intermediate
supports, if the member is continuous.
Engineering judgment must be exercised in determining whether any present flexural cracking is moderate to severe. Crack widths in reinforced concrete bridges exceeding 0.006 inches
to 0.012inches reflect the lower bound of moderate cracking. The American Concrete Institute Committee Report 224R-01 (1) presents guidance for what could be considered reasonable
or tolerable crack widths at the tensile face of reinforced concrete structures for typical conditions. These range from 0.006 inches for marine or seawater spray environments to 0.007 inches
for structures exposed to de-icing chemicals, to 0.012 inches for structures in a humid, moist
environment.
In prestressed concrete bridge structural elements, tolerable crack width criteria have
been adopted in the PCI MNL-37-06 Manual for the Evaluation and Repair of Precast Prestressed Concrete Bridge Products (20). The PCI Bridge Committee recommends that flexural
cracks greater in width than 0.006 inches should be evaluated to affirm adequate design and
performance.
Note that this attribute is a companion to the screening attribute S.4 Flexural Cracking, in
which any moderate to severe flexural cracking should exclude the bridge from a risk-based
assessment unless appropriate engineering analysis has been completed showing that the cracking is benign or has been repaired. Generally, cracking in prestressed elements is more problematic than cracking in reinforced concrete elements.

92 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Crack widths equal to or less than 0.006 inches


to 0.012 inches, depending on environment for
reinforced concrete; crack widths equal to or
less than 0.006 inches for prestressed concrete

10 points

No flexural cracking

0 points

C.15 Shear Cracking


Reason(s) for Attribute. Similar to flexural cracking, if the primary load-bearing members
in a concrete bridge exhibit shear cracking, it can be assumed that the members were either
inadequately designed for the required loading or that deterioration has occurred, which has
reduced the load-bearing capacity of the members. In either case, large shear cracks can be
indicative of an inadequate load-bearing capacity, which may require an engineering analysis
in order to determine the cause of the cracking and the resulting effect on the load capacity. As
a result, bridges exhibiting moderate to severe shear cracking should be screened from the reliability assessment unless appropriate engineering analysis indicates that the cracking is benign
in terms of the load-bearing capacity.
Assessment Procedure. Engineering judgment must be exercised in determining the severity of any present shear cracking. Shear cracks will typically present themselves with a roughly
45 degree diagonal orientation and will radiate towards the mid-span of the member for conventionally reinforced concrete. For prestressed concrete, angles down to roughly 30 degrees may
be observed. The ends of the member and any sections located over piers should be checked for
this type of cracking. Note that this attribute is a companion to the screening attribute S.5 Shear
Cracking, where any moderate to severe flexural cracking should exclude the bridge from a riskbased assessment until adequate assessments have been conducted.
Minor, hairline to less than 0.0625 inch
shear cracking

10 points

No shear cracking

0 points

C.16 Longitudinal Cracking in Prestressed Elements


Reason(s) for Attribute. This attribute is for the assessment of prestressed concrete bridge
elements. Longitudinal cracking in prestressed elements can be indicative of the corrosion or the
fracture of the embedded prestressing strands. As a result, elements with reported longitudinal
cracking in the soffit, web, or flange should be individually assessed to determine the source of
the cracking and to assess the condition of the prestressing strands (21).
Assessment Procedure. Longitudinal cracking in prestressed elements can be indicative of
strand corrosion and damage, and, as such, significant longitudinal cracking is a screening attribute. The use of longitudinal cracking in prestressed elements as a condition attribute assumes
the cracking in question is minor in nature, and significant strand corrosion is not currently
present. In this case, the longitudinal cracking provides pathways for the intrusion of moisture
and chlorides to the prestressing strands and the mild steel bars. As a result, a prestressed element
with minor longitudinal cracking is more likely to experience deterioration and damage than an
uncracked element. This attribute is scored based on inspection results.
Minor longitudinal cracking in beam soffit

15 points

No longitudinal cracking in beam soffit

0 points

Attribute Index and Commentary 93

C.17 Coating Condition


Reason(s) for Attribute. This attribute considers the effect of the coating condition
on the likelihood of corrosion damage occurring in steel bridge elements. Coatings are
applied to steel elements to provide protection from corrosion and for aesthetic reasons.
Elements with coatings in good condition, and performing as intended, are generally less
susceptible to corrosion damage. Elements with significant rusting and corrosion in areas in
which that paint system has failed are more likely to experience further corrosion damage
in the future.
Assessment Procedure. Depending on the condition of the coating, the likelihood of corrosion damage varies. Coatings typically deteriorate more rapidly where drainage from the bridge
deck is allowed to flow onto the steel surface. As a result, conditions for the accelerated corrosion
of steel may already exist. If the coating is already in poor condition, the likelihood of severe
corrosion damage is greater than for a coating in good condition. If the element is constructed
with weathering steel (assuming it is placed in the proper environment and is detailed correctly),
it should be scored as though the coating is in good condition. The development of an effective
patina for the weathering steel should be confirmed.
Coating system in very poor condition,
limited or no effectiveness for corrosion
protection, greater than 3% rusting

10 points

Coating system is in poor condition, 1% to


3% rusting, substantially effective for
corrosion protection

5 points

Coating is in fair to good condition, effective


for corrosion protection

0 points

C.18 Condition of Fatigue Cracks


Reason(s) for Attribute. Active fatigue cracks due to primary stress ranges will continue
to grow until the failure of the member, either by brittle or by ductile fracture. An arrested or
repaired fatigue crack is better than having an active crack, but it is still worse than having no
crack at all, as it suggests that the conditions necessary for cracking to initiate were or still may
be present in the structure. In other words, other similar details (that have not been preemptively
retrofitted) may be susceptible to cracking in the future.
Assessment Procedure. To determine whether or not a fatigue crack is arrested, a comparison
must be made between previous inspection reports. In order to be considered arrested, a crack
must have not grown in a specified amount of time (e.g., the inspection interval plus one year). It
is noted that although no fatigue cracks may have been observed, a detail still may be highly susceptible to fatigue. Hence, other attributes such as D.16 Element Connection Type, D.17 Worst
Fatigue Detail Category, and L.1 ADTT are included in the assessment procedure to address the
susceptibility to cracking.
Fatigue crack exists and is active/unknown

20 points (see S.7)

Fatigue crack exists and has arrested


or been retrofitted

10 points

No fatigue cracks are present

0 points

94 Proposed Guideline for Reliability-Based Bridge Inspection Practices

C.19 Presence of Fatigue Cracks Due to Secondary or Out-of-Plane Stress


Reason(s) for Attribute. Fatigue cracks due to secondary or out-of-plane stresses are the
most common type of fatigue cracks found on highway bridges. Most of these cracks occur due
to incompatibility or relative movement between bridge components.
Assessment Procedure. The scoring for this attribute is based on the existence or non
existence of fatigue cracks. Some common types of fatigue cracks due to secondary stresses
include web-gap cracks, deck plate cracking in orthotropic bridge decks, and floor beam
connections.
Fatigue cracks are present and are active/unknown

15 points

Fatigue cracks are present but have been arrested


or have been retrofitted

5 points

No fatigue cracks are present

0 points

C.20 Non-Fatigue-Related Cracks or Defects


Reason(s) for Attribute. This attribute refers to steel bridge elements that may be susceptible to fatigue-induced cracking. Fatigue cracks generally start from some initial crack or defect.
As a result of this, fatigue and brittle fracture is less likely if there are no cracks or defects from
which cracks can propagate.
Assessment Procedure. This attribute should be scored based on whether or not cracks or
other defects are found in the element. Previous inspection reports should be used when evaluating this attribute.
Non-fatigue-related cracks or defects are present

10 points

Non-fatigue-related cracks or defects are


not present

0 points

C.21 Presence of Active Corrosion


Reason(s) for Attribute. The presence of visible active corrosion on steel bridge elements
indicates that severe corrosion damage in the future is possible, since the environment and
the bridge features are vulnerable to the initiation and the propagation of corrosion. It is also
well known that corrosion damage typically propagates at an accelerated rate, once initiated,
and that elements that show no signs of active corrosion are very unlikely to develop severe
corrosion damage during the assessment interval of 72 months. Maximum rates of section loss
under the most severe marine conditions typically do not exceed 10 mils/year (0.010 inches/
year). For moderate conditions, rates are typically on the order of 4 mils/year (0.004 inches/
year) or less.
Corrosion damage that is inactive is explicitly distinguished from corrosion that is active.
For example, section loss on a girder web that was the result of a leaking expansion joint that
was corrected (the joint was replaced and the girder was repainted), may be assumed to have
inactive corrosion. It is assumed that the owner has determined that the existing section loss is
either insignificant or has taken it into account in the rating procedures and that load posting,
if needed, is in place.
Assessment Procedure. This attribute should be scored based on the amount of active corrosion present on the element. Engineering judgment should be used in determining whether

Attribute Index and Commentary 95

or not the corrosion is active. This attribute may also be used as a screening tool in a reliability
assessment.
Significant amount of active corrosion present

20 points

Moderate amount of active corrosion present

15 points

Minor amount of active corrosion present

7 points

No active corrosion present

0 points

C.22 Presence of Debris


Reason(s) for Attribute. The presence of debris on bridge elements can substantially
increase the probability of corrosion damage by maintaining a moisture-rich environment on
the surface of the steel. Debris can be especially damaging if it is allowed to remain on the bridge
without maintenance action, such as washing or cleaning. This attribute is intended to characterize bridges susceptible to having debris deposited on the flanges, bearings, connections, or
other details that results in atypical (e.g., accelerated) deterioration patterns.
Assessment Procedure. This attribute should be assessed based on if debris is present or
likely to be present on the element, resulting in an atypical deterioration pattern.
Debris is or is likely to be present

15 points

Debris not likely to be present

0 points

References
1. ACI Committee 224, ACI 224R-01: Control of Cracking in Concrete Structures. 2001, American Concrete
Institute: Farmington Hill, Michigan.
2. Fisher, J.W., and Lichtenstein, A., Hoan Bridge Forensic Investigation Failure Analysis Final Report. 2001,
Wisconsin Department of Transportation: Madison, WI.
3. FHWA, Recording and Coding Guide for the Structure Inventory and Appraisal of the Nations Bridges. 1995,
Federal Highway Administration: Washington, D.C.
4. Skeet, J., G. Kriviak, and M. Chichak, Service Life Prediction of Protective System for Concrete Bridge Decks in
Alberta. 1994, Edmonton, Alberta, Canada: Alberta Transportation and Utilities, Research & Development.
5. Russell, H.G., NCHRP Synthesis 333: Concrete Bridge Deck Performance. 2004, Transportation Research
Board of the National Academies, Washington, D.C.
6. Clemena, G.G. and Y.P. Virmani, Corrosion Protection: Concrete Bridges. 1998, McLean, VA: U.S. Dept. of
Transportation, Federal Highway Administration, Research and Development, Turner-Fairbank Highway
Research Center.
7. FHWA, Uncoated Weathering Steel in Structures, in Federal Highway Administration Technical Advisory
T-5140.22. 1989, FHWA: Washington, D.C.
8. Huang, H., Shenton, H.W., and Chajes, M.J. Load distribution for a highly skewed bridge: Testing and
analysis. Journal of Bridge Engineering, 2004, 9(6), 558562: ASCE, Reston, VA.
9. Coletti, D., B. Chavel, and W.J. Gatti, Challenges of Skew in Bridges with Steel Girders. Transportation
Research Record: Journal of the Transportation Research Board, No. 2251 2011: Transportation Research Board
of the National Academies, Washington, D.C., pp. 4756.
10. Fu, G., Feng, J., Dimaria, J., and Zhuang, Y., Bridge Deck Corner Cracking on Skewed Structures, 2007. MDOT
Report RC 1490.
11. Menassa, C., Mabsout, M., Tarhini, K., and Frederick, G. Influence of Skew Angle on Reinforced Concrete Slab
Bridges. Journal of Bridge Engineering, 2007. 12(2), pp. 205214: ASCE, Reston, VA.
12. Tindal, T.T. and Yoo, C.H. Thermal Effects on Skewed Steel Highway Bridges and Bearing Orientation.
Journal of Bridge Engineering, 2003, 8(2), p. 5765: ASCE, Reston, VA.
13. Connor, R.J. and M.J. Parr, A Method for Determining the Interval for Hands-On Inspection of Steel Bridges
with Fracture Critical Members. 2008: Purdue University. p. 32.

96 Proposed Guideline for Reliability-Based Bridge Inspection Practices

14. McLean, D.I., et al., NCHRP Synthesis 266: Dynamic Impact Factors for Bridges,. 1998, TRB, National
Research Council: Washington, D.C.
15. Silano, L.G. and P. Brinckerhoff, Bridge Inspection and Rehabilitation: A Practical Guide. 1993, New York, NY:
John Wiley & Sons, Inc.
16. Connor, R.J., M.R. Urban, and E.J. Kaufmann, NCHRP Report 604: Heat Straightening Repair of Damaged Steel Bridge GirdersFatigue and Fracture Performance,. 2008, Transportation Research Board of the
National Academies, Washington, D.C.
17. Weyers, R.E., et al., SHRP-S-360: Bridge Protection, Repair, and Rehabilitation Relative to Reinforcement
Corrosion: A Methods Application Manual, Strategic Highway Research Program, report. 1993, Transportation
Research Board of the National Academies: Washington, D.C.
18. FHWA, FHWA-RD-99-177: Portland Cement Concrete (PCC) Partial-Depth Spall Repair, 1999, Federal
Highway Administration: McLean, VA.
19. Oak Ridge, N.L., Primer on Durability of Nuclear Power Plant Reinforced Concrete StructuresA Review of
Pertinent Factors. 2006, U.S. Nuclear Regulatory Commission. p. 114.
20. PCI, Manual for the Evaluation and Repair of Precast, Prestressed Concrete Bridge Products: Including Imperfections or Damage Occurring During Production, Handling, Transportation, and Erection. 2006: Chicago, IL.
21. Naito, C., Sause, R., Hodgson, I., Pessiki, S., and Macioce, T., Forensic Examination of a Noncomposite Adjacent Precast Prestressed Concrete Box Beam Bridge. Journal of Bridge Engineering, 2010, 15(4), p. 408418:
ASCE, Reston, VA.

APPENDIX F

Illustrative Examples
98
98
98
98
99
99
99
99
100
102
105
106
107
107

109
109
109
109
110
110
110
110
112
112
114
115
116
117

117
117
117
117
118
118
119
119
121
123
124
126
126

F 1Introduction
F 2 Example 1: Prestressed Concrete Bridge
F 2.1 Bridge Profile
F 2.1.1 Overview
F 2.1.2 Concrete Bridge Deck
F 2.1.3 Prestressed Girders
F 2.1.4 Substructure
F 2.2 Assessment
F 2.2.1 Concrete Bridge Deck
F 2.2.2 Prestressed Girder
F 2.2.3 Substructure
F 2.3 Consequence Assessment
F 2.4 Scoring Summary
F 2.5 Criteria for a Family of Bridges

F 3 Example 2: Steel Girder Bridge


F 3.1 Bridge Profile
F 3.1.1 Overview
F 3.1.2 Concrete Bridge Deck
F 3.1.3 Steel Girders
F 3.1.4 Substructure
F 3.2 Assessment
F 3.2.1 Concrete Bridge Deck
F 3.2.2 Asphalt Overlay
F 3.2.3 Steel Girders
F 3.2.4 Substructure
F 3.3 Consequence Assessment
F 3.4 Scoring Summary
F 3.5 Inspection Data

F 4 Example 3: Reinforced Concrete Bridge


F 4.1 Bridge Profile
F 4.1.1 Overview
F 4.1.2 Concrete Bridge Deck
F 4.1.3 Reinforced Concrete Girders
F 4.1.4 Substructure
F 4.2 Assessment
F 4.2.1 Concrete Bridge Deck
F 4.2.2 Reinforced Concrete Girders
F 4.2.3 Substructure
F 4.3 Consequence
F 4.4 Scoring Summary
F 4.5 Inspection Data
97

98 Proposed Guideline for Reliability-Based Bridge Inspection Practices

F 1 Introduction
This section provides three illustrative examples of applying reliability-based analysis to
establish an inspection interval and strategy. The first is an example of a bridge constructed
with a superstructure composed of prestressed girders, the second example is a bridge
with a multi-girder steel superstructure, and the third example is a multi-girder reinforced
concrete superstructure. The RAP assembled by a bridge owner would typically conduct this
analysis. For these examples, typical attributes that could be identified by a RAP have been
selected for illustrative purposes. Attribute scoring sheets are shown to illustrate the process
of applying a numerical scoring process for identified attributes to estimate the reliability
of bridge elements, and to develop rationale for determining the appropriate inspection
interval.
In the examples shown, Occurrence Factor (OF) categories were determined by applying the
following equation:
X=

Si 4
So

Where Si is the score recorded for each attribute and So is the maximum score for each
attribute, such that the ratio Si So is a value between 0 and 1. OFs were then applied
such that values of X between 0 and <1 were identified as Remote, values 1 or greater but
less than 2 Low, etc. This provides a simple methodology for ranking bridges according
to their important attributes that contribute to the durability and reliability of the bridge,
and estimating the appropriate OF. This scoring methodology should be calibrated by the
RAP for its specific bridge inventory to ensure results are consistent with sound engineering
judgment.
The examples also describe the Consequence Factors that were selected for each bridge, along
with the rationale for selection. Based on these results, an appropriate inspection interval is
identified for each bridge based on the risk matrix (Figure C1). The IPN for each damage mode
is also calculated to illustrate how the process prioritizes damage modes to support inspection
procedures for that bridge.

F 2 Example 1: Prestressed Concrete Bridge


F 2.1 Bridge Profile
F 2.1.1 Overview
This example bridge is constructed of prestressed girders with a composite concrete
deck (Figure F1). The bridge has a typical reinforced concrete deck, seven prestressed AASHTO
Type IV girders, and a reinforced concrete substructure. The bridge was constructed in 2006.
Epoxy-coated reinforcement has been used in the deck and in parts of the prestressed
girders. The substructure contains regular, uncoated reinforcement. The rate of de-icing
chemical application is moderate, and the environment is also moderate. The reported
ADTT is 210 vehicles. An element-level inspection had been conducted on the bridge,
and data from the element-level inspection including inspector notes were used in deter
mining values for the condition attributes. All elements were rated 100% in Condition
State (CS) 1.

Illustrative Examples 99

Figure F1. Elevation view of Example Bridge 1.

F 2.1.2 Concrete Bridge Deck


The deck for this structure was cast-in-place and constructed with normal concrete and
epoxy-coated rebar. From the design plans, the concrete cover for the top of the deck is 1- inches.
Asphaltic plug joints in the deck are in good condition.
Some transverse cracks, spaced 2 to 3 feet apart, have been noted on the underside of the deck.
Efflorescence is present near these cracks, though there is no rust staining. No other damage has
been observed. The current condition rating is 7-Good Condition, based on the most recent
inspection.
F 2.1.3 Prestressed Girders
The superstructure of this bridge consists of 7 AASHTO Type IV prestressed concrete girders.
There is at least 2 inches of clear cover for all surfaces as determined from the design plans, and
the mild reinforcing is epoxy coated. No sealers or coatings have been applied to the girders. The
maximum span length is 99 feet. The superstructure has no observed spalling or cracking and
was most recently rated as being condition 8-Very Good Condition.
F 2.1.4 Substructure
The substructure was constructed of normal concrete with uncoated carbon steel reinforcement. The minimum design cover was determined to be 2 inches. Water from the deck does not
contact the substructure either through the drainage system or through the joints. There are no
observed signs of cracking or spalling. No evidence of unusual rotation or settlement has been
noted, and the bridge is founded on rock. The substructure is rated to have a condition rating of
8-Very Good Condition based on the most recent inspection report.

F 2.2 Assessment
This section will show how the methodology is applied to determine the OFs, the
Consequence Factors, and the corresponding inspection intervals for this bridge. A detailed
scoring of each damage mode will be presented with written descriptions of how the

100 Proposed Guideline for Reliability-Based Bridge Inspection Practices

consequence of damage was considered. The results are then summarized in a table that
provides the maximum inspection interval based on the risk matrix and the IPN determined
from the analysis.
The primary elements of this bridge are a concrete bridge deck, prestressed concrete girders,
piers, and abutments. For the concrete bridge deck element, the RAP identified typical damage
modes of widespread corrosion-induced cracking and spalling. Since each of these damage mode
results from the effects of corrosion, these damage modes were combined into a single damage
mode named Corrosion Damage.
For the prestressed concrete girders, the RAP identified the following damage modes:



Bearing Area Damage,


Corrosion Between Beam Ends,
Flexural and Shear Cracking, and
Strand Fracture.
For the substructure, the damage mode considered was:

Corrosion Damage (cracking and spalling due to the effects of corrosion).

Considering the damage modes identified for each element, attributes relating to each
damage mode were identified and ranked, as described in the Guideline. The following sections contain illustrative examples of attribute scoring sheets developed for the different
elements and damage modes for the bridge and the estimated OFs based on the attribute
scoring.
F 2.2.1 Concrete Bridge Deck
The RAP determined that certain attributes of a bridge deck that contribute to the likelihood of
corrosion damage are common and well known, and that these same attributes would generally
apply to other bridge decks in its inventory, as well as other typical concrete elements. Additionally, because corrosion will affect most concrete elements and associated damage modes, repetition of certain common attributes could be reduced by having a single corrosion profile for an
element. This corrosion profile could then be applied to all damage modes stemming from corrosion for a given element more efficiently. As such, a corrosion profile was developed to assess
the corrosion-resistance characteristics of a concrete bridge deck or other concrete element. This
profile included typical attributes that were well known to affect the durability of concrete, but did
not depend on the current condition or individual characteristics of an element. The attributes
identified included:








Poor Deck Drainage and Ponding,


Years of Construction,
Application of Protective Systems,
Concrete Mix Design,
Minimum Concrete Cover,
Reinforcement Type,
Exposure Environment,
Rate of De-icing Chemical Application, and
Maintenance Cycle.

Supporting rationale for each of these attributes from the commentary (Appendix E) was
used. Utilizing these corrosion profile attributes and the suggested rankings in the commentary,
the RAP developed a simple scoring sheet to calculate the corrosion profile for a bridge deck as
shown in the table below.

Illustrative Examples 101

Corrosion Profile, Concrete Bridge Deck


Attribute

Score

D.4 Poor Deck Drainage and Ponding


The deck drainage system is of modern design and is effective

D.6 Year of Construction


Bridge constructed in 2006

D.7 Application of Protective Systems


Protective systems never applied to deck

10

D.8 Concrete Mix Design


Constructed of normal grade concrete, no admixtures

15

D.11 Minimum Concrete Cover


Design cover is 1.5 inches

10

D.12 Reinforcement Type


Epoxy-coated reinforcement used

L.3 Exposure Environment


Deck environment is moderate

10

L.5 Rate of De-icing Chemical Application


Rate of de-icing chemical application is moderate

15

C.5 Maintenance Cycle


Bridge receives regular, periodic maintenance

Corrosion Profile score

60 out of 140

Attributes were identified by the RAP that affected the reliability and durability of a bare
concrete deck. These attributes include the corrosion profile score, plus attributes based on the
loading and the condition of a particular deck. The RAP identified screening criteria of the Current Condition Rating and Fire Damage for concrete bridge decks, to identify decks that may
require further assessment. Other attributes of bare concrete decks were identified and ranked.
The scoring plan was then applied to the subject concrete deck.

Corrosion Damage, Concrete Bridge Deck


Attribute

Score

S.1 Current Condition Rating


Current deck condition rating is greater than 4

Pass

S.2 Fire Damage


No fire damage in the past 12 months

Pass

Corrosion Profile score

60

L.1 ADTT
ADTT is moderate (210 vehicles)

10

C.1 Current Condition Rating


Current deck condition rating is 7

C.8 Corrosion-Induced Cracking


Minor corrosion-induced cracking noted

102 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Corrosion Damage, Concrete Bridge Deck


Attribute
C.9 General Cracking
No general cracking observed

Score
0

C.10 Delaminations
No delaminations found

C.11 Presence of Repaired Areas


No repaired areas

C.12 Presence of Spalling


No spalling noted

C.13 Efflorescence/Staining
Minor efflorescence without rust observed

Corrosion Damage total


Corrosion Damage ranking

80 out of 290
1.1 Low

This bridge deck is still relatively new, was built to modern standards for durability and corrosion resistance, and has very little damage accumulation. As a result, the deck received very
low scores for the attributes identified. Based on the attribute score, the RAP estimated that the
likelihood of the failure for the deck (based on the criteria described in Section 2.1) in the next
72 months was low, i.e., the OF was Low (OF = 2).
F 2.2.2 Prestressed Girder
For the assessment of a prestressed girder, the corrosion profile scoring model was also used.
As with the corrosion profile for bridge decks, this basic profile can be applied across many
concrete elements. In this case, the prestressed girder scored the same as the deck.
Corrosion Profile, Prestressed Girder
Attribute

Score

D.4 Poor Deck Drainage and Ponding


The deck drainage system is of modern design and is effective

D.6 Year of Construction


Bridge constructed in 2006

D.7 Application of Protective Systems


Protective systems never applied

10

D.8 Concrete Mix Design


Constructed of normal grade concrete

15

D.11 Minimum Concrete Cover


Minimum concrete cover is 2 inches

10

D.12 Reinforcement Type


Reinforcement is epoxy coated

L.3 Exposure Environment


Superstructure environment is moderate

10

L.5 Rate of De-icing Chemical Application


Rate of salt application is moderate

15

Illustrative Examples 103

Corrosion Profile, Prestressed Girder


Attribute
C.5 Maintenance Cycle
Bridge receives regular, periodic maintenance
Corrosion Profile point total

Score
0
60 out of 140

The RAP then considered the identified damage modes for a prestressed girder element, identified and ranked attributes, and applied the scoring model for each damage mode as shown
below.
Bearing Area Damage, Prestressed Girder
Attribute

Score

Corrosion Profile score

60

D.1 Joint Type


Bridge contains a closed joint system

C.4 Joint Condition


Joints are not leaking

C.8 Corrosion-Induced Cracking


No corrosion-induced cracking noted

C.9 General Cracking


No general cracking observed

C.11 Presence of Repaired Areas


No repaired areas

C.12 Presence of Spalling


No areas of spalling noted

Bearing Area Damage point total


Bearing Area Damage ranking

65 out of 240
1.08 Low

Corrosion Between Beam Ends, Prestressed Girder


Attribute
Corrosion Profile score

Score
60

C.8 Corrosion-Induced Cracking


No corrosion-induced cracking noted

C.10 Delaminations
No delaminations found

C.11 Presence of Repaired Areas


No repaired areas

C.12 Presence of Spalling


No spalling present

C.13 Efflorescence/Staining
No signs of efflorescence

104 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Corrosion Between Beam Ends, Prestressed Girder


Attribute
Corrosion Between Beam Ends point total
Corrosion Between Beam Ends ranking
Flexural/Shear Cracking, Prestressed Girder
Attribute
S.4 Flexural Cracking
No flexural cracking
S.5 Shear Cracking
No shear cracking
D.2 Load Posting
Bridge is not load posted
L.4 Likelihood of Overload
Likelihood of overload is low
C.14 Flexural Cracking
No flexural cracking
C.15 Shear Cracking
No shear cracking
Flexural/Shear Cracking point total
Flexural/Shear Cracking ranking
Strand Fracture, Prestressed Girder
Attribute
S.1 Current Condition Rating
Superstructure condition rating is greater than 4
S.6 Longitudinal Cracking in Prestressed Elements
Significant cracking is not present
Corrosion Profile score
L.6 Subjected to Overspray
Bridge not over a roadway, not exposed to overspray
C.1 Current Condition Rating
Superstructure condition rating is 8
C.4 Joint Condition
Joints are present but not leaking
C.8 Corrosion-Induced Cracking
No corrosion-induced cracking noted
C.10 Delaminations
No delaminations found
C.11 Presence of Repaired Areas
No repaired areas
C.12 Presence of Spalling
No spalling present

Score
60 out of 235
1.02 Low

Score
Pass
Pass
0
0
0
0
0 out of 55
0 Remote

Score
Pass
Pass
60
0
0
5
0
0
0
0

Illustrative Examples 105

Strand Fracture, Prestressed Girder


Attribute
C.16 Longitudinal Cracking in Prestressed Elements
No longitudinal cracking in the girders
Strand Fracture point total
Strand Fracture ranking

Score
0
65 out of 285
0.91 Remote

Based on the attributes identified by the RAP, the OF for the bearing area damage and corrosion between the beam ends was estimated to be Low (OF = 2). For the damage modes of shear
cracking, flexural cracking and strand fracture, the OF was Remote (OF = 1).
F 2.2.3 Substructure
For the piers and abutments, the RAP considered that the most likely damage modes were
corrosion-induced cracking and spalling, or a settlement or rotation of one of the substructure elements. However, settlement and rotations were determined to not be relevant damage
modes because the bridge substructure is founded on rock. To estimate the likelihood for the
corrosion damage mode, the panel once again used the generalized corrosion profile scoring.
The panel then considered appropriate attributes for estimating the OF for the corrosion
damage mode, identified and ranked key attributes, and scored the piers and abutments for
the bridge, as shown below.

Corrosion Profile, Substructure


Attribute

Score

D.4 Poor Deck Drainage and Ponding


Deck does not drain onto the substructure

D.6 Year of Construction


Bridge constructed in 2006

D.7 Application of Protective Systems


Protective systems never applied

10

D.8 Concrete Mix Design


Substructure constructed with normal grade concrete

15

D.11 Minimum Concrete Cover


Minimum design cover is 2 inches

10

D.12 Reinforcement Type


Reinforcement is uncoated carbon steel

15

L.3 Exposure Environment


Environment is rated as moderate

10

L.5 Rate of De-icing Chemical Application


Rate of de-icing chemical application is moderate

15

C.5 Maintenance Cycle


Bridge receives regular, periodic maintenance

Corrosion Profile point total

75 out of 140

106 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Corrosion DamagePiers and Abutments, Substructure


Attribute

Score

Corrosion Profile score

75

C.1 Current Condition Rating


Current substructure condition rating is 8

C.4 Joint Condition


Joints present but not leaking

C.8 Corrosion-Induced Cracking


No corrosion-induced cracking noted

C.9 General Cracking


No cracking observed

C.10 Delaminations
No delaminations found

C.11 Presence of Repaired Areas


No repaired areas present

C.12 Presence of Spalling


No spalling noted

C.13 Efflorescence/Staining
No signs of efflorescence

Corrosion Damage point total

80 out of 290

Corrosion Damage ranking

1.10 Low

Based on the attribute scoring, the OF for the damage mode of Corrosion Damage was
assessed to be Low (OF = 2).

F 2.3 Consequence Assessment


Once the likelihood for each damage mode has been ranked, the RAP must perform a consequence analysis for each damage mode considered. For the concrete bridge deck, based on
the damage mode of corrosion damage, the RAP considered the scenario of significant spalling
of the deck as a result of extensive corrosion damage. Since the bridge is over a non-navigable
waterway, spalling of concrete from the soffit would have a low consequence. Considering the
ADT and the posted speed limit, spalling on the deck surface was determined to have only a
moderate effect on serviceability for the bridge and a planned repair. The consensus of the RAP
was that the appropriate Consequence Factor was Moderate (CF = 2). The RAPs consequence
assessment will be included in the file for the bridge.
For the prestressed girder superstructure, in order to determine the consequence of failure,
the RAP considered the scenario that one of the prestressed beams lost 100% of its load carrying
capacity due to the damage modes of strand fracture, flexural and shear cracking, or corrosion
between the beam ends. The RAP reviewed data from two very similar bridges for which truck
impacts severely damaged one or more of the prestressed girders. The RAP determined that these
two bridges could be considered very similar as their span lengths were within 10% of the bridge
under consideration, and had nearly identical girder spacing and deck configuration. In both cases,
the impact severely damaged at least one of the girders such that its load carrying capacity was
effectively reduced to 0. The bridges exhibited little or no additional dead load deflection and were
capable of carrying normal live loads. Temporary barriers were installed to shift traffic away from
the shoulder area above the fascia girders that were damaged. Further, the load rating information

Illustrative Examples 107

for this bridge was reviewed and the bridge possessed a capacity far in excess of the required Inventory and Operating ratings. Hence, the RAP concluded that the loss of one girder would at most
have a Moderate (CF = 2) consequence based on the following rationale:
The bridge is redundant, based on AASHTO definitions;
The bridge is very similar to other bridges for which a member failure has occurred, but did

not result in collapse of the bridge or excessive deflection;


The bridge capacity far exceeds required Inventory and Operating ratings;
The bridge has low ADT, such that there will not be a major impact on traffic; and
The bridge is located over a non-navigable stream. Thus, the risks to people or property under

the bridge are minimal.


For the damage mode of bearing area damage, two scenarios were considered. The first scenario
considered that the bearing area damage was sufficient to result in a downward displacement of the
bridge deck. The most likely consequence was assessed by the panel to be Moderate, because such
a displacement would result in only moderate disruption of service and require a planned repair.
This was based on the rationale that the deck is composite with the superstructure and the bridge is
a multi-girder bridge with normal beam spacing, such that any displacement would be minor and
localized in nature, because loads could transfer to adjacent girders and the composite deck would
limit displacements. The second scenario considered was that the bearing area damage resulted in
severe cracking in the shear area of the beam, resulting in damage to the development length of
the strands or shear cracking. The RAP considered that such a scenario would, at worst, result in
100% loss in load carrying capacity, as was considered for the damage modes of strand fracture,
flexural or shear cracking, and corrosion between the beams ends in the previous scenario. Based
on these two scenarios, the CF of Moderate (CF = 2) was selected for this damage mode. The RAPs
consequence assessment will be included in the file for the bridge.
For the reinforced concrete substructure, the RAP considered the scenario that there was widespread corrosion damage (cracking and spalling) to the piers and abutments. The bridge is over a
small creek, and hence there is little concern of injury from spalling concrete. The piers and abutments are short. Past experience of the panel with many piers and abutments of similar characteristics indicated that serious corrosion damage has a benign immediate effect on serviceability and
safety. Therefore, the consensus of the panel was that the appropriate consequence category was
Low (CF = 1).
The data from the RAP assessment was then applied to the appropriate risk matrix (Figure C1)
to determine the maximum inspection interval for the bridge. A summary of the scoring and maximum inspection interval for the bridge are shown below.

F 2.4 Scoring Summary


Table F1 shows a summary of the analysis for this bridge. The maximum inspection interval based
on the RAP analysis was determined to be 72 months, based on the low likelihood of serious damage
(failure) to the elements of the bridge, and the moderate consequences associated with that damage.

F 2.5 Criteria for a Family of Bridges


The RAP assessed that it has many bridges in its inventory of very similar design characteristics. Based on the key attributes developed by the RAP, the panel identified a series of criteria to
apply to a family of bridges to extend this analysis to other bridges in its inventory. These criteria
describe bridges of the same design type and characteristics, with similarly adequate load ratings,
and similar environmental loading. Condition attributes were mapped to suitable surrogates in the
element-level bridge inspection data that were being collected for the bridge. For example, for the
prestressed concrete girders, the panel identified that the individual condition attributes identified

108 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table F1. Reliability assessment scoring summary for Example Bridge 1.
Element
Deck
Prestressed
Girders

Substructure

Damage
Corrosion
Damage
Bearing Area
Damage
Corrosion
Between Beam
Ends
Flexural/Shear
Cracking
Strand Fracture
Corrosion
Damage

Occurrence
Factor (OF)

Consequence
Factor (CF)

Maximum
Interval

OF x CF
(IPN)

Low (2)

Moderate (2)

72 months

Low (2)

Moderate (2)

72 months

Low (2)

Moderate (2)

72 months

Remote (1)

Moderate (2)

72 months

Remote (1)

Moderate (2)

72 months

Low (2)

Low (1)

72 months

by the analysis, such as shear or flexural cracking, corrosion-induced cracking, spalling, or efflorescence, were either not present or minimal if the CS ratings for the element were CS 1 or CS 2. Therefore, for the prestressed girder element, elements that are rated as CS 1 and CS 2 would not have the
damage characteristics the panel identified as key to the potential for serious damage to develop.
Bridges with any portion of the prestressed element rated as CS 3 would likely have one or more of
these condition attributes present, and therefore would require reanalysis and possibly a reduced
inspection interval. Similar criteria were developed for each of the elements assessed by the RAP.
The RAP also identified that longitudinal cracking in prestressed elements was a key condition attribute not adequately represented in its element-level inspection scheme. As a result,
the RBI procedure for bridges in this family needed to include a requirement that longitudinal
cracking be assessed during the inspection. This requirement was included in the RBI procedure
as a special emphasis area for this family of bridges.
The RAP developed a listing of criteria, including design characteristics and using surrogate
element data for certain condition attributes, to apply to the overall family of similar bridges in
its inventory. These criteria are based on the engineering assessment documented through the
RAP analysis. Example criteria to identify the family of bridges included:









Maximum span length less than 120 feet;


Four or more AASHTO prestressed girders;
Beam spacing of 10 feet or less;
ADTT less than 1000;
Constructed in 1995 or later;
No structural element with CS 3 reported;
No joint element with CS 3 reported;
Load rating exceeds requirements;
No significant flexural, shear, or longitudinal cracking in the prestressed element; and
Bridge receives RBI-based inspections.

The RAP determined that bridges meeting these criteria will be treated as a family under
the RBI methodology. If a particular bridge violates any of these criteria, it must be reassessed
according to the attribute scoring criteria developed for this family of bridges.
Table F2 summarizes the information from the RAP analysis to be included in the RBI procedure for these bridges. Longitudinal cracking in the prestressed elements is indicated as a special
emphasis area for the inspection, to ensure this key damage mode is assessed during subsequent
inspections. Other IPNs for identified damage modes are low, indicating a standard RBI inspection is required for the bridge.

Illustrative Examples 109


Table F2. Table of information to be included in the
RBI procedure.
Maximum Inspection Interval: 72 months
Special Emphasis Items
S.6 Longitudinal Cracking in Prestressed Elements

Element
Deck
Prestressed Girder

Substructure

RBI Damage Modes


Damage Mode
Corrosion Damage
Bearing Area Damage
Corrosion Between Beam Ends
Flexural/Shear Cracking
Strand Fracture
Corrosion Damage

IPN
4
4
4
2
2
2

F 3 Example 2: Steel Girder Bridge


F 3.1 Bridge Profile
F 3.1.1 Overview
This example bridge carries a state highway over a non-navigable river. The bridge was constructed in 1954 with a continuous steel girder superstructure, a non-composite reinforced concrete deck, and a reinforced concrete substructure (Figure F2). All steel reinforcement used in
this bridge is regular uncoated mild carbon steel. The observed ADTT is 130 vehicles. The rate of
salt application is determined to be high by the RAP, with more than 100 applications of de-icing
chemicals per year. The exposure environment is considered moderate.
F 3.1.2 Concrete Bridge Deck
The reinforced concrete bridge deck was constructed of cast-in-place normal concrete. From
the design plans, the minimum cover was determined to be 1-916 inches. The deck has a bituminous wearing surface of unknown thickness which was assessed to be in fair condition. In some
locations the wearing surface has come off the deck. No membranes or sealers have been applied.
The deck has no reported drainage or ponding problems.

Figure F2. Elevation view of Example Bridge 2.

110 Proposed Guideline for Reliability-Based Bridge Inspection Practices

The most recent inspection rated the deck condition as 6-Satisfactory. According to the
inspection report, the underside of the deck has hairline transverse cracks, spaced 2 to 3 feet
apart, with efflorescence stains. The underside of the approach span at abutment 1 has heavy
efflorescence stains on the left side.
F 3.1.3 Steel Girders
The continuous steel girder superstructure is constructed from four painted steel girders with
steel diaphragms. These girders are riveted at the connection plates. No problems were found
at the connection plates during a recent in-depth inspection. The bottom flanges of the girders
have corrosion with missing paint. These locations have some pack rust formation. The superstructure was assessed to have a condition rating of 6-Satisfactory.
Based on the inspection report, no fatigue or fracture related damage is present. Based on the
provided design plans, it was determined that the girders are riveted built-up members, so the
worst fatigue detail category is D.
F 3.1.4 Substructure
The substructure was constructed of normal grade reinforced concrete with uncoated carbon
steel reinforcement. The minimum cover was determined to be 3-38 inches. Drainage from the
deck is leaking onto the substructure from the deck due to leaking joints.
There is no observed evidence of rotation or settlement. The concrete piers have random hairline cracks with some moderate surface scaling below the high water line. Hairline to 132 inch
(0.03125 inch) diagonal and vertical cracks with minor efflorescence stains have been reported
on the concrete abutments. The concrete pier caps have some hairline cracks but appear to be in
good condition. There is spalling in the concrete piers exposing rebar. The substructure condition
was assessed to be 6-Satisfactory.

F 3.2 Assessment
The primary elements of this bridge are a concrete bridge deck with an asphalt overlay, riveted
steel girders, deck joints, piers, and abutments. For the concrete bridge deck element the typical
damage modes identified were concrete cracking and spalling. Since each of these damage modes
results from the effects of corrosion, these damage modes were again grouped into a single damage mode termed Corrosion Damage. The same corrosion profile as developed for the previous example was used for the deck. The asphalt overlay for the deck was assessed individually
for debonding and spalling/potholes. For the steel girders, the damage modes considered were:
Corrosion Damage,
Fatigue Damage, and
Fracture Damage.

For the substructure, the damage mode considered was:


Corrosion Damage (cracking and spalling due to the effects of corrosion).

The RAP determined through consensus that tilting of the piers or unexpected settlement
were not credible damage modes. This was based on the rationale that the bridge had been in
service for more than 50 years without any signs of tilt or rotation, the geographic area was not
susceptible to subsurface erosion or unexpected settlements, and the roller bearings were insensitive to moderate displacements of the substructure.
F 3.2.1 Concrete Bridge Deck
The concrete deck was assessed for the damage mode of corrosion damage, using the corrosion
profile for concrete elements and attributes identified for the deck, as shown below.

Illustrative Examples 111

Corrosion Profile, Concrete Bridge Deck


Attribute

Score

D.4 Poor Deck Drainage and Ponding


No drainage problems noted

D.6 Year of Construction


Bridge constructed in 1954

D.7 Application of Protective Systems


Protective systems never applied to deck

10

D.8 Concrete Mix Design


Constructed of normal grade concrete, no admixtures

15

D.11 Minimum Concrete Cover


Design cover is between 1.5 inches and 2.5 inches

10

D.12 Reinforcement Type


Uncoated carbon steel reinforcement

15

L.3 Exposure Environment


Deck environment is moderate

10

L.5 Rate of De-icing Chemical Application


Rate of de-icing chemical application is high (100 times per year)

20

C.5 Maintenance Cycle


Maintenance cycle is at least limited

10

Corrosion Profile score

96 out of 140

Corrosion Damage, Concrete Bridge Deck


Attribute

Score

S.1 Current Condition Rating


Current deck condition rating is greater than 4

Pass

S.2 Fire Damage


No fire damage in the past 12 months

Pass

Corrosion Profile score

96

L.1 ADTT
ADTT is minor (130 vehicles)

C.1 Current Condition Rating


Current deck condition rating is 6

C.8 Corrosion-Induced Cracking


Minor corrosion-induced cracking noted

C.9 General Cracking


No general cracking observed

C.10 Delaminations
UnknownAsphalt overlay prevents effective sounding

20

C.11 Presence of Repaired Areas


No repaired areas

C.12 Presence of Spalling


No spalling noted

112 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Corrosion Damage, Concrete Bridge Deck


Attribute
C.13 Efflorescence/Staining
Moderate efflorescence without rust observed
Extent of Damage total
Corrosion damage ranking

Score
10
141 out of 290
1.94 Low

Based on the attributes identified by the RAP, the OF for corrosion damage was assessed to
be Low (OF = 2).
F 3.2.2 Asphalt Overlay
The asphalt overlay was assessed by the panel using a simple expert elicitation. The general
consensus of the panel was that the typical service life of an asphalt overlay was less than 10 years.
The RAP agreed that the likelihood of failure of the asphalt overlay was greater than 1% over a
72-month interval, given that the overlay was already in service. The OF for the overlay failure
was determined to be High (OF = 4) by consensus of the panel.
F 3.2.3 Steel Girders
The steel girders were assessed for three damage modes: Fatigue Damage, Corrosion Damage,
and Fracture Damage. Key attributes were identified by the RAP as shown below. Supporting
data and rationale for each attribute are included in the commentary.

Fatigue Damage, Steel Girder


Attribute

Score

S.7 Active Fatigue Cracks due to Primary Stress Ranges


No active fatigue cracks due to primary stress

Pass

D.6 Year of Construction


Bridge was built in 1954

20

D.16 Element Connection Type


Element is connected by rivets

D.17 Worst Fatigue Detail Category


Worst fatigue detail category is D

15

L.1 ADTT
ADTT is 130 vehicles

15

L.7 Remaining Fatigue Life


Remaining fatigue life is unknown

10

C.18 Condition of Fatigue Cracks


No fatigue cracks present

C.19 Presence of Fatigue Cracks due to Secondary or Out-of-Plane Stress


No fatigue cracks due to secondary or out of plane stress

Fatigue Damage point total


Fatigue Damage ranking

67 out of 110
2.44 Moderate

Illustrative Examples 113

Corrosion Damage, Steel Girder


Attribute

Score

S.9 Significant Level of Active Corrosion or Section Loss


Active corrosion present is not alarming

Pass

D.5 Use of Open Decking


Bridge does not have an open deck

D.13 Built-Up Member


Element is built up

15

D.15 Constructed of Weathering Steel


Element not constructed with weathering steel

10

L.3 Exposure Environment


Exposure environment is moderate

10

L.5 Rate of De-icing Chemical Application


Rate of de-icing chemical application is high (100 times per year)

20

L.6 Subjected to Overspray


Superstructure is not subjected to overspray

C.4 Joint Condition


Joints are moderately leaking

15

C.7 Quality of Deck Drainage System


Drainage system is of adequate quality

C.17 Coating Condition


Element is painted, with steel exposed on bottom flanges

10

C.21 Presence of Active Corrosion


Significant active corrosion is present

20

C.22 Presence of Debris


Element has no debris

Corrosion Damage point total

100 out of 190

Corrosion Damage ranking

2.1 Moderate

Fracture Damage, Steel Girder


Attribute

Score

S.7 Active Fatigue Cracks due to Primary Stress Ranges


No active fatigue cracks due to primary stress

Pass

S.8 Details Susceptible to Constraint-Induced Fracture


No details susceptible to constraint induced fracture

Pass

D.3 Minimum Vertical Clearance


Bridge is not over a roadway, max vertical clearance

D.6 Year of Construction


Bridge constructed in 1954

20

D.14 Constructed of High Performance Steel


Element is not constructed of HPS/unknown

10

L.1 ADTT
ADTT is 130 vehicles

15

114 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Fracture Damage, Steel Girder


Attribute
L.7 Remaining Fatigue Life
Remaining fatigue life is unknown

Score
10

C.6 Previously Impacted


Bridge has not been impacted before

C.19 Presence of Fatigue Cracks due to Secondary or Out-of-Plane Stress


No fatigue cracks present

C.20 Non-Fatigue-Related Cracks or Defects


No fatigue cracks present

Fracture Damage point total


Fracture Damage ranking

55 out of 125
1.76 Low

The RAP analysis of key attributes for the damage modes indicated that the steel superstructure has a moderate likelihood of fatigue damage (OF = 3), a moderate likelihood of developing
corrosion damage (OF = 3), and a low likelihood of fracture (OF = 2).
F 3.2.4 Substructure
The substructure was assessed for the damage mode of corrosion damage, using the corrosion
profile for concrete elements and attributes identified for the piers and abutments.

Corrosion Profile, Substructure


Attribute

Score

D.4 Poor Deck Drainage and Ponding


No drainage problems noted

D.6 Year of Construction


Bridge constructed in 1954

D.7 Application of Protective Systems


Protective systems have not been applied

10

D.8 Concrete Mix Design


Substructure constructed with normal grade concrete, no admixtures

15

D.11 Minimum Concrete Cover


Minimum design concrete cover is 3-38

D.12 Reinforcement Type


Reinforcement is uncoated carbon steel

15

L.3 Exposure Environment


Exposure environment is moderate

10

L.5 Rate of De-icing Chemical Application


Rate of de-icing chemical application is high (100 times per year)

20

C.5 Maintenance Cycle


Maintenance cycle is at least limited

10

Corrosion Profile point total

86 out of 140

Illustrative Examples 115

Corrosion DamagePiers and Abutments, Substructure


Attribute

Score

Corrosion Profile score

86

C.1 Current Condition Rating


Current substructure condition rating is six

C.4 Joint Condition


Joints are significantly leaking onto substructure

20

C.8 Corrosion-Induced Cracking


Moderate corrosion-induced cracking noted

10

C.9 General Cracking


Presence of minor general cracking

C.10 Delaminations
Minor localized delaminations on footings

C.11 Presence of Repaired Areas


No repaired areas present

C.12 Presence of Spalling


Significant spalling with exposed reinforcement present on piers

20

C.13 Efflorescence/Staining
Moderate efflorescence without rust staining

10

Substructure Elements point total

161 out of 290

Substructure Elements ranking

2.22 Moderate

Based on the attribute scoring, the RAP estimated the OF was Moderate (OF = 3) for corrosion
damage for the piers and abutments. A considerable amount of damage has already accumulated
in the form of spalling with exposed reinforcement and moderate cracking.

F 3.3 Consequence Assessment


Since the bridge carries a state highway over a non-navigable river, key damage to the bridge
deck is likely to be in the form of spalling on the riding surface of the bridge deck. The most
likely consequence of severe damage to the deck is Moderate (CF = 2) because there may be some
disruption of service or reduction in posted speed. The bridge is a four girder bridge with typical
girder spacing, such that even a through-thickness punch-through is likely to be local in nature
and not represent a high consequence. The assignment of a moderate consequence is based on
common experience with bridge decks of similar design characteristics.
The consequence of the asphalt overlay failing was determined to be Low, because failure of
the asphalt overlay was a maintenance need and would not necessitate increased inspection or
monitoring.
The superstructure consists of four steel girders with diaphragms spaced at 20 to 25 feet.
Although fatigue damage is the most likely damage mode, the worst outcome associated with
fatigue would be the fracture of one of the girders. Hence, the consequence scenario evaluated was
the fracture of one of the girders. Note that this analysis does not depend on the damage failure
mode, thus, failure could also be due to corrosion. As stated, the cross section is made up of four
identical built-up members. In evaluating the most likely consequence, the RAP identified several
similar designs where full-depth fractures of steel girders occurred. These bridges had spans greater

116 Proposed Guideline for Reliability-Based Bridge Inspection Practices

than or equal to this bridge, had similar skew, had similar girder spacing, and had a non-composite
deck. In all cases, none of the bridges collapsed, though some displayed minor sagging. The bridges
carried full service load up until the time that fracture was detected in later inspections. Hence, the
RAP determined that the consequence associated with fracture of one of the girders should be set
as High (CF = 3) based on the following rationale:
The bridge is redundant, based on AASHTO definitions;
The bridge is very similar to other bridges where full-depth girder fractures occurred, but did

not result in collapse of the bridge or excessive deflection;


The bridge meets required Inventory and Operating ratings;
Fracture in a member will have a major impact on travel, since the member failure would

result in a lane closure; and


The bridge is located over a non-navigable river. Thus, the risks to people or property under
the bridge are minimal.
The RAPs consequence assessment will be included in the bridge file along with appropriate
references to the other bridges cited in the consequence scenario evaluation.
Engineering calculations showing that the effects of a girder fracture would result in a Moderate consequence (CF = 2) would be required to reduce the consequence category for this scenario. Based on the above and the fact that the bridge is not fracture-critical, the consequence
category of Severe was not considered a plausible outcome for girder fracture.
For the substructure, the scenario considered for damage to the piers and abutments of the
bridge was severe corrosion damage and spalling. The most likely consequence of this scenario
is a Low consequence (OF = 1), because severe corrosion damage of this type would typically
require monitoring and assessment, but would not affect the serviceability of the bridge.
The summary of the RAP assessment is shown in Table F3. Based on this assessment, the maximum inspection interval for this bridge is 24 months, due to the likelihood and high consequence
associated with the development of fatigue cracking. This is due to in part to the fact that the bridge
has fatigue-prone details (category D), the bridge was constructed before modern fracture control
requirements were in place, and there is truck traffic on the bridge. Even though the bridge has not
developed any fatigue cracks in more than 50 years of service, the rational assessment performed by
the RAP indicates that the potential for cracking exists, and should be treated appropriately. Additionally, the bridge is susceptible to serious corrosion damage, because its current condition includes
active corrosion, the applications of de-icing chemical are high, the members are built up, and the
joints are leaking. As such, the required maximum interval for an RBI is 24 months.

F 3.4 Scoring Summary


The scoring summary for this bridge is shown in Table F3. Based on the reliability assessment,
the maximum inspection interval was determined to be 24 months.
Table F3. Reliability assessment scoring summary for Example Bridge 2.
Element

Damage

Deck
Overlay
Steel Girders

Corrosion Damage
Debonding/Spalling
Fatigue
Corrosion
Fracture
Corrosion Damage

Substructure

Occurrence
Factor (OF)
Low (2)
High (4)
Moderate (3)
Moderate (3)
Low (2)
Moderate (3)

Consequence
Factor (CF)
Moderate (2)
Low (1)
High (3)
High (3)
High (3)
Low (1)

Interval
72 months
48 months
24 months
24 months
48 months
48 months

OF x CF
(IPN)
4
4
9
9
6
3

Illustrative Examples 117


Table F4. Table of information to be included in the RBI.
Maximum Inspection Interval: 24 Months
Special Emphasis Items
S.7 Active Fatigue Cracks due to Primary Stress Ranges
S.9 Significant Level of Active Corrosion or Section Loss
RBI Damage Modes
Element
Damage Mode
Corrosion Damage
Deck
Fatigue Cracking
Steel Girder
Corrosion Damage
Fracture
Corrosion Damage
Substructure

IPN
4
9
9
6
3

F 3.5 Inspection Data


Table F4 summarizes the information from the RAP analysis to be included in the RBI procedure to be used for this bridge. The identified screening criteria for fatigue cracking due to primary stresses and significant section loss are included as special emphasis items. The data in the
table also indicates that fatigue cracking and corrosion damage are priority items for inspection
of the steel girders, based on their IPN of 9. Because of the high IPN for corrosion damage, the
RAP recommends utilizing an ultrasonic thickness gauge (UT-T) to assess the areas of section
loss in the steel girder. This will ensure accurate reporting of the remaining section and mitigate
the risks associated with severe section loss, which was identified through the RAP analysis as
being moderately likely to occur, and resulting in a high consequence. For fatigue cracking, the
high IPN number will prioritize fatigue cracking for the inspection team conducting the RBI on
the bridge. The IPN of 9 indicates to the inspector that the bridge has the potential for fatigue
cracking, and the consequences of that cracking are potentially high were it to go undetected.
The RAP could recommend NDE, such as Magnetic Particle Testing (MT) or Dye Penetrant
Testing (PT), be applied to a sampling of locations during periodic inspections to ensure that
fatigue cracking is detected and enhance the reliability of the inspection.

F 4 Example 3: Reinforced Concrete Bridge


F 4.1 Bridge Profile
F 4.1.1 Overview
This example bridge is a typical, simply-supported three-span reinforced concrete bridge with
a bare cast-in-place deck. The bridge owners inventory includes more than 100 bridges of similar span length and design characteristics, and, as such, is developing the RAP analysis for application to a family of bridges, using this bridge as an example of the family. The specific bridge
was constructed in 1963 and carries highway traffic over a local road. The estimated ADT on the
bridge is 22,000 vehicles, while the ADT on the local road under the bridge is 60 vehicles. Both
the rate of salt application and the surrounding environment are considered to be moderate. A
photograph of the bridge is shown in Figure F3.
F 4.1.2 Concrete Bridge Deck
For this bridge, the deck was constructed with normal grade cast-in-place concrete and
uncoated mild steel reinforcement. The asphalt has been removed from the top of the deck and
a water proof sealant has been applied. Hairline to 116-inch cracks have been observed on the
top of the deck near the abutments. Hairline diagonal cracks with efflorescence stains have been
observed on the soffit of the deck near the abutments. No delaminations or spalling are noted
on the deck.

118 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Figure F3. Elevation view of Example Bridge 3.

From the design plans, the minimum cover was determined to be 1-1316 inches. Based on the
most recent inspection report, the deck is considered to be in CS 6-Satisfactory. This deck contains concrete edge joints with silicon sealant. The seals are considered to be in good condition
but are leaking water. No other ponding or drainage issues are noted.
F 4.1.3 Reinforced Concrete Girders
The superstructure for this bridge consists of seven reinforced concrete girders that
are constructed from normal grade concrete and uncoated mild steel reinforcement. Each
girder, per span, has hairline vertical flexure cracking. The right exterior girder has a spall
on the bottom end which measures 12 inches tall by 3 inches wide by 5 inches deep due
to impact.
One of the exterior girders has an 8-inch diameter spall resulting from an over-height vehicle
collision. Girders five and six also have scrapes and spalls from an over-height vehicle collision.
The superstructure is considered to be in CS 5-Fair. From the design plans, the minimum concrete cover is 3-58 inches.
F 4.1.4 Substructure
The substructure for this bridge is also constructed of normal grade concrete with
uncoated mild steel reinforcement. From the design plans, the minimum cover was determined to be 2- inches. The columns have random hairline cracks and the top of column
four has an area of delamination that is 29 inches tall by 21 inches wide. Both abutments
have hairline to 116-inch vertical cracks and spalling with exposed reinforcement on their
right sides.
All bents have water staining resulting from leaking joints. Bent cap one, span one, has horizontal cracks with delamination in the bottom left corner. Bent cap two, span two, has an area
of cracking and delamination that is 16 inches wide by 8 inches tall near girder six. Bent cap two,
span three, also has an area of cracking and delamination that is 27 inches wide by 4 inches tall
near girder six.
The substructure has neoprene pad bearings which have curled on the ends but are still in
satisfactory condition. The overall condition rating for the substructure is 5-Fair. There are no
signs of settlement or rotation and the substructure itself is founded on rock.

Illustrative Examples 119

F 4.2 Assessment
The primary elements of this bridge are a concrete bridge deck, reinforced concrete girders, and piers and abutments. For the concrete bridge deck element, the typical damage mode
identified was corrosion damage (concrete cracking and spalling). The same corrosion profile
developed for the previous examples was also used for this deck. For the reinforced concrete
girders, the damage modes considered were:
Bearing Area Damage,
Corrosion Between Beam Ends, and
Flexural and Shear Cracking.

Based on the owners inventory data and experience, there has been no occurrences of significant
shear cracking in bridges of similar design to the one being analyzed. However, there have been
isolated cases of cracking due to flexural stresses, possibly resulting from overloaded trucks. Based
on this experience, the RAP determines that flexural cracking is an important damage mode, while
the likelihood of shear cracking is more remote, generally. To provide focus on the flexural cracking
experience in this particular inventory, the RAP determines that shear cracking and flexural cracking should be separated into distinct damage modes. Additionally, the RAP determined through
consensus that the likelihood of overload would have the greatest influence on the likelihood of
flexural cracking progressing; existing flexural cracking had moderate effect, and the fact that bridge
may be load posted has only a small effect. As such the RAP assigns 20 points to L.4, Likelihood of
Overload, only 10 points to D.2, Load Posting and 15 points to C.14, Flexural Cracking. The key
attributes for flexural cracking were therefore determined by the RAP to be as follows:



S.4 Flexural Cracking (screening criteria),


D.2 Load Posting,
L.4 Likelihood of Overload, and
C.14 Flexural Cracking.

The screening criteria for Flexural Cracking (S.4) was also utilized to identify bridges with
significant flexural cracking, which may require individual engineering assessment. For shear
cracking, the relevant attributes identified by the RAP were:



S.5 Shear Cracking (screening),


D.2 Load Posting,
L.4 Likelihood of Overload, and
C.15 Shear Cracking.

Again, the screening attribute S.5 for unresolved shear cracking is utilized to identify any
bridges with shear cracking that may require engineering assessment.
For the substructure, the damage mode considered was:
Corrosion Damage (cracking and spalling due to the effects of corrosion).

F 4.2.1 Concrete Bridge Deck


The concrete deck was assessed for the damage mode of corrosion damage, using the corrosion profile for concrete elements and attributes identified for the deck, as shown below

Corrosion Profile, Concrete Bridge Deck


Attribute
D.4 Poor Deck Drainage and Ponding
No drainage problems noted

Score
0

120 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Corrosion Profile, Concrete Bridge Deck


Attribute
D.6 Year of Construction
Bridge constructed in 1963

Score
6

D.7 Application of Protective Systems


Waterproof penetrating sealer applied, frequency unknown

D.8 Concrete Mix Design


Constructed of normal grade concrete, no admixtures

15

D.11 Minimum Concrete Cover


Design cover is between 1.5 inches and 2.5 inches

10

D.12 Reinforcement Type


Uncoated carbon steel reinforcement

15

L.3 Exposure Environment


Deck environment is moderate

10

L.5 Rate of De-icing Chemical Application


Rate of de-icing chemical application is moderate

15

C.5 Maintenance Cycle


Maintenance cycle is at least limited

10

Corrosion Profile score

86 out of 140

Corrosion Damage, Concrete Bridge Deck


Attribute

Score

S.1 Current Condition Rating


Current deck condition rating is greater than four

Pass

S.2 Fire Damage


No fire damage in the past 12 months

Pass

Corrosion Profile score

86

L.1 ADTT
ADTT is high (5,500 vehicles)

20

C.1 Current Condition Rating


Current deck condition rating is 6

C.8 Corrosion-Induced Cracking


Moderate corrosion-induced cracking noted

10

C.9 General Cracking


Moderate general cracking observed

10

C.10 Delaminations
No delaminations noted

C.11 Presence of Repaired Areas


No repaired areas

C.12 Presence of Spalling


No spalling noted

Illustrative Examples 121

Corrosion Damage, Concrete Bridge Deck


Attribute
C.13 Efflorescence/Staining
Minor efflorescence without rust observed
Corrosion Damage total
Corrosion Damage ranking

Score
5
136 out of 290
1.88 Low

Based on the attributes identified by the RAP, the OF for corrosion damage in the deck was
estimated as Low (OF = 2).
F 4.2.2 Reinforced Concrete Girders
The reinforced concrete girders were assessed for the damage modes of bearing area damage,
corrosion between the beam ends, and flexural and shear cracking.
Corrosion Profile, Reinforced Concrete Girder
Attribute

Score

D.4 Poor Deck Drainage and Ponding


No drainage problems noted.

D.6 Year of Construction


Bridge constructed in 1963

D.7 Application of Protective Systems


Protective systems never applied

10

D.8 Concrete Mix Design


Constructed of normal grade concrete

15

D.11 Minimum Concrete Cover


Minimum concrete cover is greater than 2.5 inches

D.12 Reinforcement Type


Reinforcement is uncoated mild steel

15

L.3 Exposure Environment


Superstructure environment is moderate

10

L.5 Rate of De-icing Chemical Application


Rate of salt application is moderate

15

C.5 Maintenance Cycle


Bridge maintenance is at least limited

10

Corrosion Profile point total

81 out of 140

Bearing Area Damage, Reinforced Concrete Girder


Attribute

Score

Corrosion Profile score

81

D.1 Joint Type


Bridge has closed joints

122 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Bearing Area Damage, Reinforced Concrete Girder


Attribute
C.4 Joint Condition
Joints are leaking but sealant is still in fair condition

Score
15

C.8 Corrosion-Induced Cracking


No corrosion-induced cracking noted

C.9 General Cracking


No general cracking observed

C.11 Presence of Repaired Areas


No repaired areas

C.12 Presence of Spalling


Moderate spalling in several locations, no exposed reinforcement noted.

15

Bearing Area Damage point total


Bearing Area Damage ranking

111 out of 240


1.85 Low

Corrosion Between Beam Ends, Reinforced Concrete Girder


Attribute

Score

Corrosion Profile score

81

C.1 Current Condition Rating


Current condition rating is 5

20

C.6 Previously Impacted

20

C.8 Corrosion-Induced Cracking


No corrosion-induced cracking noted

C.9 General Cracking


No general cracking observed

C.10 Delaminations
Unknown

20

C.11 Presence of Repaired Areas


No repaired areas

C.12 Presence of Spalling


Moderate spalling in several locations (due to impact), no exposed
reinforcement noted
C.13 Efflorescence/Staining
No signs of efflorescence

15

Corrosion Between Beam Ends point total

156 out of 290

Corrosion Between Beam Ends ranking

2.15 Moderate

Flexural Cracking, Reinforced Concrete Girder


Attribute

Score

S.4 Flexural Cracking


Hairline flexural cracking noted, determined to be benign

Pass

Illustrative Examples 123

Flexural Cracking, Reinforced Concrete Girder


Attribute
D.2 Load Posting
Bridge is not load posted

Score
0

L.4 Likelihood of Overload


Likelihood of overload is moderate

10

C.14 Flexural Cracking


Hairline flexural cracking noted

15

Flexural Cracking point total


Flexural Cracking ranking

25 out of 45
2.22 Moderate

Shear Cracking, Reinforced Concrete Girder


Attribute

Score

S.5 Shear Cracking


No shear cracking present

Pass

D.2 Load Posting


Bridge is not load posted

L.4 Likelihood of Overload


Likelihood of overload is moderate

10

C.15 Shear Cracking


No shear cracking

Shear Cracking point total

10 out of 45

Shear Cracking ranking

0.88 Remote

The attribute scoring indicated an OF of Moderate (OF = 3) for corrosion between beam ends
and flexural cracking, an OF of Low (OF = 2) for bearing area damage, and an OF of Remote
(OF = 1) for shear cracking.
F 4.2.3 Substructure
The substructure was assessed for the damage mode of corrosion damage, using the corrosion
profile for concrete elements and attributes identified for the piers and abutments.
Corrosion Profile, Substructure
Attribute

Score

D.4 Poor Deck Drainage and Ponding


No drainage problems noted

D.6 Year of Construction


Bridge constructed in 1963

D.7 Application of Protective Systems


Protective systems have not been applied

10

D.8 Concrete Mix Design


Substructure constructed with normal grade concrete, no admixtures

15

124 Proposed Guideline for Reliability-Based Bridge Inspection Practices

Corrosion Profile, Substructure


Attribute
D.11 Minimum Concrete Cover
Minimum design concrete cover is 2- inches

Score
0

D.12 Reinforcement Type


Reinforcement is uncoated carbon steel

15

L.3 Exposure Environment


Exposure environment is moderate

10

L.5 Rate of De-icing Chemical Application


Rate of de-icing chemical application is moderate

15

C.5 Maintenance Cycle


Maintenance cycle is at least limited

10

Corrosion Profile point total

81 out of 140

Corrosion DamagePiers and Abutments, Substructure


Attribute

Score

Corrosion Profile score

81

C.1 Current Condition Rating


Current substructure condition rating is five

20

C.4 Joint Condition


Joints are moderately leaking onto substructure

15

C.8 Corrosion-Induced Cracking


Localized cracking near delaminations noted

C.9 General Cracking


Presence of moderate general cracking

10

C.10 Delaminations
Unknown

20

C.11 Presence of Repaired Areas


No repaired areas present

C.12 Presence of Spalling


Moderate spalling with exposed reinforcement present

15

C.13 Efflorescence/Staining
No efflorescence noted

Concrete Elements point total

166 out of 290

Concrete Elements ranking

2.28 Moderate

Based on their analysis, the RAP assessed that the likelihood of failure due to corrosion damage
was moderate for the pier and abutments (OF = 3). Already, a considerable amount of damage has
accumulated in the form of localized delaminations and spalling resulting in exposed reinforcement.

F 4.3 Consequence
For the concrete bridge deck, the RAP considered the scenario that the corrosion damage
in the deck resulted in spalling of either the driving surface of the deck or deck soffit. In this

Illustrative Examples 125

case, the bridge carries a high-volume highway over another, lower-volume roadway. The
roadway on the bridge carries 22,000 vehicles a day, and the roadway below the bridge carries
60 vehicles a day. Based on this information, any spalling from the deck soffit has the potential
to fall into the roadway below and strike a motorist. However, given the low traffic volume
and speed on the roadway below, the RAP considered the likelihood of this occurring to be
relatively small. Therefore, the consensus of the RAP was that the appropriate Consequence
Factor was High (CF = 3). For spalling of the riding surface, the panel determined that such a
scenario was likely to have an effect on serviceability of the deck, and may require a reduction
in the posted traffic speed. Therefore, the consensus of the RAP was that this represented a
Consequence Factor of Moderate (CF = 2). For this case, the scenario of concrete falling into
the roadway below the bridge provides the Consequence Factor for corrosion damage in the
bridge deck.
To determine the Consequence Factor for the concrete beams, the RAP considered the
scenario that one of the reinforced concrete beams lost 100% of its load carrying capacity due to corrosion damage between the beam ends, flexural or shear cracking, or bearing
area damage. The RAP considered that the superstructure is reinforced concrete with a
composite deck such that redundancy in the structure would prevent the total collapse
of a girder. The RAP also reviewed data from two very similar bridges for which corrosion damage had resulted in loss of load carrying capacity in one girder of a multi-girder,
reinforced concrete bridge with a composite deck. The RAP determined that these two
bridges could be considered very similar to the bridge being analyzed because their span
lengths were within 10% of the bridge under consideration and they utilized a nearly identical girder spacing and deck configuration. In both cases, the corrosion damage had reduced
a single girders load carrying capacity effectively to zero, however, the bridge exhibited little
or no additional dead load deflection and was capable of carrying normal live loads. Lane closures were required on the bridges as the result of the faulted girder, resulting in a significant
impact on traffic.
The load rating information for the bridge was reviewed and the bridge possessed a capacity
far in excess of the required Inventory and Operating ratings. However, the bridge carries a high
ADT, such that a lane closure would have a major impact on traffic. Additionally, the roadway
under the bridge is a low-volume road that may be impacted by the shoring required or debris.
As a result, the Consequences Factor was determined to be High (CF = 3) based on the following
rationale:
The bridge is redundant, based on AASHTO definitions;
The bridge is very similar to other bridges for which a member failure has occurred, but did

not result in collapse of the bridge or excessive deflection;


The bridge capacity far exceeds required Inventory and Operating ratings;
The bridge has high ADT, such that there could be a major impact on traffic; and
The bridge is located over a low-volume roadway such that there would be some risks to traffic
on the roadway below.
For the damage mode of bearing area damage, two scenarios were considered. The first
scenario considered that the bearing area damage was sufficient to result in a downward displacement of the bridge deck. The most likely consequence was assessed by the panel to be
Moderate (CF = 2), resulting in only a minor disruption of service, since the deck is composite
with the superstructure and it is a multi-girder bridge with normal beam spacing. The second
scenario considered was that the bearing area damage resulted in severe cracking in the shear
area of the beam, resulting in a loss of load-carry capacity. As above, the Consequence Factor
of High (CF = 3) was assigned.

126 Proposed Guideline for Reliability-Based Bridge Inspection Practices


Table F5. Reliability assessment scoring summary for Example Bridge 3.
Element

Damage

Deck
Reinforced
Concrete
Girders

Corrosion Damage
Bearing Area
Damage
Corrosion
Between Beam
Ends
Flexural Cracking
Shear Cracking

Substructure

Corrosion Damage

Occurrence
Factor (OF)
Low (2)

Consequence
Factor (CF)
High (3)

Maximum
Interval
48 months

OF x CF
(IPN)
6

Low (2)

High (3)

48 months

Moderate (3)

High (3)

24 months

Moderate (3)
Remote (1)

High (3)
High (3)

24 months
48 Months

9
3

Moderate (3)

High (3)

24 months

For the reinforced concrete substructure, areas of delaminations are present in several locations and both abutments have areas of spalling with exposed reinforcement. Here, the most
likely damage mode will result in spalling of the concrete. The RAP considers that this bridge is
located over a roadway, and the piers are immediately adjacent to the roadway such there is a
chance that concrete spalling off of a pier could strike a passing motorist. Based on this factor,
the consequence scenario for this damage mode was assessed to be High (CF = 3).

F 4.4 Scoring Summary


Table F5 shows a summary of the scoring for this bridge. Based on the likelihood of corrosion damage between the beam ends, flexural cracking and corrosion damage to the substructure, and the associated consequences, the maximum inspection interval was determined to
be 24 months.

F 4.5 Inspection Data


Table F6 summarizes the information from the RAP analysis to be included in the RBI procedure for this bridge. This information includes the identified screening criteria of inspection for
flexural cracking as a special emphasis item for the RBI inspection.
Based on the RAP assessment, this particular bridge has high IPNs for corrosion between the
beam ends (9), corrosion damage in the substructure (9), and corrosion damage in the deck (6),

Table F6. Table of information to be included in RBI procedure.


Maximum Inspection Interval: 24 Months
Special Emphasis Items
S.4 Flexural Cracking

Element
Deck
Reinforced Concrete
Girder

Substructure

RBI Damage Modes


Damage Mode
Corrosion Damage
Bearing Area Damage
Corrosion Damage between the beam
ends
Flexural Cracking
Shear Cracking
Corrosion Damage

IPN
6
6
9
9
3
9

Illustrative Examples 127

each of which have the potential to result in debris falling into the roadway below the bridge.
As a result, the RAP determined that enhanced inspection for corrosion damage was needed as
part of the RBI procedure. Available technologies to complete the delamination survey include
hammer sounding, infrared thermography (IR) and Impact Echo (IE). The RAP recommends
delamination surveys be completed during the periodic inspections to mitigate the risk of debris
falling into the roadway below the bridge unexpectedly.
Flexural cracking also has a high IPN, indicating that this damage mode is of high importance
and needs to be prioritized during subsequent RBIs for the bridge. Flexural cracking is included
as a special emphasis item for subsequent inspections.

P ART I I

Final Research Report:


Developing Reliability-Based
Inspection Practices

131

SUMMARY

Final Research Report:


Developing Reliability-Based
Inspection Practices
Introduction
The National Bridge Inspection Standards (NBIS) mandate the frequency and methods
used for the safety inspection of highway bridges. The inspection frequency specified in the
National Bridge Inventory (NBI) is calendar-based and generally requires routine inspections to be conducted at a maximum interval of 24 months. The calendar-based inspection
interval applied uniformly across the bridge inventory results in the same inspection interval
for new bridges as for aging and deteriorated bridges. Such a uniform inspection practice
does not recognize that a newly constructed bridge, with improved durability characteristics
and a few years of exposure to the service environment, may be much less likely to develop
serious damage over a given time period than an older bridge that has been exposed to the
service environment for many years. As such, inspection needs may be less for the newer
bridge, and greater for the aging structure, relative to the uniform interval currently applied.
Bridges that are in benign, arid operating environments are currently inspected at the same
interval as bridges in aggressive marine environments, in which significant damage from
corrosion may develop much more rapidly, resulting in increased inspection needs. Further,
bridges that are known to possess good characteristics or details are treated the same as
those with characteristics or details known to perform poorly. Current practices make it
difficult to recognize if the same or improved safety and reliability could be achieved by
varying inspection methods and frequencies to meet the needs of a specific bridge, based
on its design, condition, and operational environment. A more rational approach to inspection planning would better match inspection requirements to inspection needs through
reliability-based analysis that considers the design, materials, condition, and operational
environment of a bridge.
As such, the goals of this project were to develop reliability-based inspection practices to
meet the goals of:
1. Improving the safety and reliability of bridges and
2. Optimizing resources for bridge inspection.
The objective of this project was to develop a suggested bridge inspection practice for
consideration for adoption by AASHTO. The practices developed through the project are
based on rational, reliability-based methods to ensure bridge safety, serviceability, and
effective use of resources.

132

Findings
Reliability theories and practices were applied through the research to develop a guideline
for Risk-Based Inspection (RBI) that provides a new approach for bridge inspection. The
methodology consists of bridge owners performing a reliability assessment of bridges within
their inventories to identify those bridges that are most in need of inspection to ensure
bridge safety, and those for which inspection needs are less. This assessment is conducted
by an expert panel at the owner level known as a Reliability Assessment Panel (RAP). The
RAP conducts a reliability-based engineering assessment of the likelihood of serious damage
resulting from common deterioration mechanisms, over a specified time period, and the likely
outcome or consequences if that damage were to occur. The reliability-based assessment can
be described by a simple, three-step process:
Step 1: What can go wrong, and how likely is it? Identify possible damage modes for the
elements of a selected bridge type. Consider design, loading, and condition characteristics
(attributes), and then categorize the likelihood of serious damage occurring into one of
four Occurrence Factors (OFs) ranging from remote (very unlikely) to high (very likely).
Step 2: What are the consequences? Assess the consequences in terms of safety and serviceability, assuming the given damage modes occur. Categorize the potential consequences
into one of four Consequence Factors (CFs) ranging from low (minor effect on serviceability) through severe (i.e., bridge collapse, loss of life).
Step 3: Determine the inspection interval and scope. Prioritize inspection needs and assign
an inspection interval for the bridge, based on the results of Steps 1 and 2.

Reliability matrix for


determining maximum
inspection intervals for
bridges.

This assessment is based on common and well-known design, loading, and condition
attributes that affect the durability characteristics of bridges. The attributes are identified
and prioritized through expert elicitation processes. A simple reliability matrix, shown in
the figure to the left, is used to identify the appropriate inspection interval for the bridge, based
on the reliability analysis. Damage modes that tend toward the upper right corner of the
matrix, meaning they are likely to occur and have high consequences if they did occur,
require shorter inspection intervals and possibly more intense or focused inspections. Damage modes that tend toward the lower left corner, meaning they are unlikely occur and/or
consequences are low if they did occur, require less frequent inspection.
Inspection intervals determined through the RBI process may be longer or shorter than
those specified by traditional uniform, calendar-based approaches, depending on needs
identified by the reliability-based engineering assessment. Inspections conducted under the
RBI process are typically more intense and thorough than traditional inspection practices,
and require condition assessment of bridge elements to meet the needs of the reliabilitybased assessment. Inspection needs are prioritized to improve the reliability of the inspection process, and bridge-specific inspection procedures can be developed based on the
reliability analysis. The methodology developed is intended for typical highway bridges of
common design characteristics.
The methodology developed through the research capitalizes on the extensive body of
knowledge and experience with in-service bridge behavior, and the common deterioration
mechanisms that cause bridges to deteriorate during their service lives. The process allows
for the integration of emerging technologies such as improved data on long-term bridge
performance and advanced modeling and analysis techniques, when available. The methodology was developed with suitable flexibility to address owner-specific needs and conditions,
while providing systematic processes and methods to support consistent application of the
technology.

133

The methodology developed through the research was tested using two case studies in
different states. During these case studies, the processes described in the Guideline for RBI
analysis were implemented using state forces to develop RBI intervals for typical highway
bridges with superstructures constructed of steel and prestressed members. The RBI intervals determined through the RBI were verified through analysis of historical records for a
sample of bridges in each state.
The reliability-based inspection practices developed through the research differ from
traditional, calendar-based approaches. The new approach to bridge inspection provides a methodology to improve the safety and reliability of bridges by focusing inspection resources where most needed. This also leads to optimized allocation of resources, as
inspection requirements are better matched to inspection needs through a reliability-based
engineering assessment.

Conclusions
This research developed inspection practices to meet the goals of (1) improving the safety
and reliability of bridges and (2) optimizing resources for bridge inspection. The goals of the
research have been achieved through the development of a new guideline document entitled
Proposed Guideline for Reliability-Based Bridge Inspection Practices, Part I of this report,
which has been developed based on the application of reliability theories. This document
meets the project objective of developing a suggested practice for consideration for adoption
by AASHTO, based on rational methods to ensure bridge safety, serviceability, and effective
use of resources. A reliability-based approach was fully developed and documented through
the Guideline. This new inspection paradigm could transform the calendar-based, uniform
inspection strategies currently implemented for bridge inspection to a new, reliability-based
approach that will better allocate inspection resources and improve the safety and reliability
of bridges.
The implementation of the Guideline developed through the research was tested by conducting case studies in two states. These studies demonstrated and verified the effectiveness
of the procedures developed in the research for identifying appropriate inspection intervals
for typical highway bridges. It was shown through these studies that the RBI practices identified appropriate inspection intervals of up to 72 months. It was concluded from these studies
that implementation of the RBI practices did not adversely affect the safety and serviceability
of the bridges analyzed in the study, based on the analysis of historical inspection records.
These studies also successfully demonstrated the implementation of the Guideline and the
procedures therein using state DOT personnel.
The results reported herein demonstrated and verified that inspection intervals of up
to 72 months were suitable for certain bridges. Such extended inspection intervals would
allow the reallocation of inspection resources toward bridges requiring more frequent and
in-depth inspections, resulting in improved safety and reliability of bridges. As such, the
project goals of developing a reliability-based bridge inspection practice that could improve
the safety and reliability of bridges, and optimizes the use of resources, were achieved
through the research.

Suggestions
The research reported herein has demonstrated the effectiveness of the RBI procedures for
determining suitable inspection intervals for typical highway bridges, and as such, broader
implementation of the technology is suggested.

134

The procedure, methods, and approach described herein can be applied for atypical
bridges as well. For example, non-redundant bridge members can be assessed using this
approach, as illustrated in previous research (60). The approach can also be applied to complex bridges, or to bridges with advanced deterioration. Analysis requirements may be more
detailed and advanced; development of such analysis may be pursued to provide a uniform
strategy for bridge inspection across the entire bridge inventory.
Finally, the back-casting procedure utilized herein may be considered for implementation
when RBI practices are to be used. Back-casting provides a means for verification of models
developed by the RAP and quality assurance of the RBI process. As such, the back-casting
procedure provides a critical tool for the implementation of RBI technology.

135

CHAPTER 1

Background

The periodic inspection of highway bridges in the United


States plays a critical role in ensuring the safety, serviceability, and reliability of bridges. Inspection processes
have developed over time to meet the requirements of the
National Bridge Inspections Standards (NBIS) (1) and to
meet the needs of individual bridge owners in terms of
managing and maintaining bridge inventories. The inspection frequency mandated by the NBIS requires the inspection
interval (maximum time period between inspections) not to
exceed 24 months. Based on certain criteria, that interval may
be extended up to 48 months with approval from the Federal
Highway Administration (FHWA) (2). Maximum inspection
intervals of less than 24 months are utilized for certain bridges
according to criteria developed by the bridge owner, typically
based on age and known deficiencies. Most bridge owners
utilize the maximum inspection interval of 24 months, as
mandated by the NBIS, for the majority of the bridges in an
inventory, and the reduced intervals for bridges with known
deficiencies. The uniform inspection interval of 24months
was specified at the origination of the National Bridge Inspection Program in 1971 based on experience, engineering judgment, and the best information available at the time.
The uniform approach provides a single maximum inspection interval for most bridges, regardless of the bridge age,
design, or environment. To date, this mandated inspection
interval has provided an adequate level of safety and reliability for the bridge inventory nationwide. However, such
a uniform inspection practice does not recognize that a
newly constructed bridge with improved durability characteristics and a few years of exposure to the service environment may be much less likely to developed serious damage
over a giventime interval than an older bridge that has
been exposed to the service environment for many years. As
such, inspection needs may be less for the newer bridge, and
greater forthe aging bridge, relative to the uniform interval
currently required. Bridges that are in benign, arid operating

environments are inspected at the same interval as bridges in


aggressive marine environments, where significant damage
from corrosion may develop much more rapidly, requiring
increased inspection to ensure that safety and serviceability is
maintained. Fracture critical members designed under modern criteria have vastly improved resistance to fatigue than
older bridges, and as such, the likelihood of fatigue damage
for modern bridges is much lower than for older bridges.
Newer bridges in general are designed to higher standards
with more durable materials such that their resistance to
loading and environmental effects is much greater than older
bridges. Current practices make it difficult to recognize if the
same or improved safety and reliability could be achieved by
varying inspection methods or frequencies to meet the needs
of a specific bridge based on its design, condition, and operational environment.
Recognizing the variability in the design, condition, and
operating environments of bridges would provide for inspection requirements that better meet the needs of individual
bridges and improves both bridge and inspection reliability. A more rational approach to inspection planning would
determine the interval and scope of an inspection according
to the condition of the bridge and the likelihood that damage
would occur. This would allow for resources to be focused
where most needed to ensure the safety and reliability. Such
inspection planning tools are highly developed in other
industries, using the principles of reliability and risk assessment to match inspection requirements to inspection needs.
These methodologies evaluate the specific characteristics of
components, such as resistance to damage modes, anticipated
deterioration mechanisms, current condition, and loading,
to evaluate the reliability of the component. Appropriate
inspection requirements are determined based on this evaluation, such that the safety and operation of the component
is maintained over its service life, and resources are allocated
efficiently.

136

As such, the goals of this project were to develop reliabilitybased inspection practices to meet the goals of:
(1) Improving the safety and reliability of bridges and
(2) Optimizing resources for bridge inspection.
The objective of this project was to develop a proposed
bridge inspection practice for consideration for adoption

by AASHTO. The practices developed through the project


are based on rational methods that ensure bridge safety,
serviceability, and effective use of resources. This report
includes an overview of the inspection planning process
that is based on the reliability principles developed during this project, and is documented in Part I of this report:
Proposed Guideline for Reliability-Based Bridge Inspection Practices.

137

CHAPTER 2

Research Approach

The Guideline were developed in consideration of modern


industrial practices and is the result of an exhaustive review
and analysis of current methodologies and practices for the
inspection and management of structures and facilities,
assessment of needs and capabilities, and the development
of methodologies focused on the unique needs of highway
bridges. The research was aimed at identifying the most effective strategy regarding development of a reliability-based
bridge inspection practice. Through this investigation, a systematic process for determining the frequency and scope of
highway bridge inspections has been developed based on reliability concepts.
Theories and practices for applying reliability concepts
are increasingly popular as a basis for design codes as a means
of adopting a more scientific basis for estimating variations
in loading and resistance (strength) of components. Applying
reliability theories in this context typically includes probabilistic analysis to deal with uncertainties in the design parameters and loading. There have been attempts to apply these
design reliability concepts to maintenance and inspection
activities, and some of this prior work will be discussed in
this report. Unfortunately, such probabilistic approaches are,
in most cases, found to be exceptionally complex and often
require assumptions regarding the future behavior and performance of bridges that are difficult to verify. Additionally,
probabilistic methods are typically focused on predicting
strength, and do not address the serviceability requirements
that are important in terms of bridge inspection. As such,
alternative methodologies were sought through the course of
the research.
In industrial applications, the more common terminology for inspection practices that use reliability theories for
development of inspection and maintenance strategies is
risk-based, with reliability being one component of a risk
analysis that also includes consideration of the consequences
of some type of failure or loss of service. Sometimes, reliability and risk terms are used interchangeably. An extensive

study of the current state-of-the-practice and state-of-theart for reliability and RBI practice was conducted as part of
this project to determine the most applicable methodologies
for the inspection of highway bridges. The best practices and
the successful implementations of these inspection practices
were reviewed, analyzed, and considered by the Research
Team. An expert panel meeting/workshop was held that
included bridge inspection experts from state departments
of transportation to provide bridge-owner perspective on the
tools being developed through the research.
Several different approaches for developing a reliabilitybased inspection practice for highway bridges were considered, ranging from pure probabilistic structural reliability
theories to fully qualitative risk analysis. The system that
was developed is intended to incorporate the best practices
and concepts from both schools of thought. The resulting
methodology provides a reliability-based inspection practice
that is implementable within the existing bridge inspection
programs in the United States. Important consideration in
developing the methodology included:
The approach should be practically implementable and

realistic.
The approach needs to be sufficiently flexible to meet the

needs of states with different inspection programs and


bridge management approaches.
The approach must be effective in ensuring bridge safety.
The approach should match inspection requirements with
inspection needs.
The approach should capitalize on the existing body of
knowledge regarding in-service bridge behavior.
Based on these considerations, a reliability-based methodology was developed for risk-based bridge inspection. In
summary, the methodology developed has its foundation
based on risk analysis that includes both the anticipated reliability of bridges (and their elements) and the consequences

138

of damage to a bridge. The methodology is strongly grounded


in existing industrial practice.
The methodology described in this report has been developed based on the well-established methods used in other
industries for practical inspection planning. Such industrial
standards, which are discussed in detail in the project interim

report (3), provide a technical foundation for the methodology developed. The approach has been customized to provide
a practical, implementable tool that can be expanded and
developed over time. The research resulted in the development of the Guideline, which documents the tools, methodologies, and requirements for RBI practices.

139

CHAPTER 3

Findings and Applications

3.1Introduction
The Guideline developed under this project describes the
methodology for RBI practices for typical highway bridges.
The goal of the methodology is to improve the safety and reliability of bridges by focusing inspection efforts where most
needed and optimizing the use of resources. The Guideline
provides a framework and procedures for developing suitable
inspection strategies, based on a rational, reliability-based
engineering assessment of inspection needs. The methodology
considers the structure type, age, condition, environment, loading, prior problems, and other characteristics that contribute to
the reliability and durability of highway bridges.
Generally, the methodology involves bridge owners performing a reliability assessment of bridges within their bridge
inventory to identify those bridges that are most in need of
inspection to ensure bridge safety, and those for which inspection needs are less. The assessment is conducted by considering the reliability and safety attributes of bridges and bridge
elements. This reliability assessment is conducted by an expert
panel assembled by a bridge owner (e.g., state) known as an
RAP. This panel conducts an engineering assessment of the
likelihood of serious damage resulting from common deterioration mechanisms, over a specified time period, applied to
key elements of a bridge. This assessment is based on common
and well-known design, loading, and condition attributes that
affect the reliability characteristics of bridge elements. These
attributes influence the likelihood that a particular element
will fail over a given time period, i.e., its reliability. The attributes are identified and prioritized through an expert elicitation process. This process capitalizes on the experience and
knowledge of bridge owners regarding the performance of
the bridges within specific operational environments, given
typical loading patterns, ambient environmental conditions,
construction quality, etc.
The reliability estimate is combined with an evaluation of
the potential outcomes or consequences, in terms of safety

and serviceability, of damage progressing to a defined failure


state. These data are then used to determine and prioritize
inspection needs for specific bridges, or families of bridges
with very similar design and condition characteristics. This
includes determining a suitable inspection interval and scope,
or procedures, to be used in the inspection. Under this process, the inspection interval is not fixed, such as it is in a uniform, calendar-based system, but rather is adjusted to meet
the anticipated needs of the specific bridge or bridges in a
family. Therefore, bridges with highly reliable characteristics,
which are unlikely to have serious deterioration over a specified time, typically have a longer inspection interval than a
bridge with less reliable characteristics, or for which the consequences of a failure may be more severe. For example, a
bridge in good condition with highly durable and redundant
design characteristics may have a longer inspection interval
than a bridge in poor condition, lacking modern durability characteristics, and/or having a non-redundant design.
Through this process, inspection resources can be focused
where most needed to ensure the safety and serviceability
of bridges. Inspection needs are prioritized to improve the
safety and reliability of the bridge inventory overall.
The approach developed under the research is a risk-based
approach that differs from purely reliability-based approaches
in that the likelihood of failure is combined explicitly with the
consequences of that failure. Risk can be defined generally as
the product of the probability of an event and the associated
consequences:
Risk = Probability Consequence
Probability in this equation is the likelihood of an adverse
event or failure occurring during a given time period. This is
sometimes expressed quantitatively as a probability of failure
(POF) estimate for a given time interval, or as a qualitative
assessment of the likelihood of an adverse event based on experience and engineering judgment. Generally, this probability is

140

the complement of the reliability. Consequence is a measure of


the impact of the event occurring, which may be measured in
terms of economic, social, safety, or environmental impacts.
The Guideline developed through this research was focused
on the inspection of typical highway bridges of common
design characteristics. Atypical structures, such as long-span
truss bridges, cable-stayed bridges, suspension bridges, and
other unique or unusual bridge designs may require certain
considerations not presently captured in the Guideline. Scour
and underwater inspections have existing methodologies for
evaluation, and as such are not included in the Guideline.
Bridges assessed using this methodology are assumed to have a
current load rating that indicates that the structural capacity is
sufficient to carry allowable loads.

3.2 Overview of Methodology


The RBI process involves an owner (e.g., state) establishing an expert panel (RAP) to define and assess the durability and reliability characteristics of bridges within that state.
The RAP uses engineering rationale, experience, and typical
deterioration patterns to evaluate the reliability characteristics of bridges and the potential outcomes of damage. This is
done through a relatively simple process that consists of three
primary steps:

from current inspection practices generally, because the damage modes typical for the specific bridge are identified and
prioritized. The inspection is required to be capable of
assessing each of these damage modes sufficiently to support
the assessment of future needs. As a result, the inspections
may be more thorough than traditional practices, including
hands-on access to key portions of a bridge such that damage is
effectively identified to support the RBI assessment. The results
of the inspection are assessed to determine if the existing RBI
procedure needs to be modified or updated as a result of findings from the inspection. For example, as a bridge deteriorates
over time and damage develops, as reported in the inspection
results, inspection intervals may be reduced to address the
inspection needs for the bridge as it ages.
The overall process for assessment under the developed
Guideline is shown schematically in Figure 1. The process
begins with the selection of a bridge or family of similar
bridges to be analyzed. For the selected bridge or bridges, the
RAP identifies common damage modes for elements of the
bridge given the design, materials, and operational environment. Key attributes are identified and ranked to assess OFs
that categorize the likelihood of serious damage developing
over a specified time interval. CFs that categorize the poten-

Step 1: What can go wrong, and how likely is it? Identify possible damage modes for the elements of a selected bridge
type. Considering design, loading, and condition characteristics (attributes), categorize the likelihood of serious
damage occurring into one of four Occurrence Factors
(OFs) ranging from remote (very unlikely) to high (very
likely).
Step 2: What are the consequences? Assess the consequences in
terms of safety and serviceability assuming the given damage modes occur. Categorize the potential consequences
into one of four Consequence Factors (CFs) ranging from
low (minor effect on serviceability) through severe (i.e.,
bridge collapse, loss of life).
Step 3: Determine the inspection interval and scope. Use
a simple reliability matrix to prioritize inspection needs
and assign an inspection interval for the bridge based on
the results of Steps 1 and 2. Damage modes that are likely
to occur and have high consequences are prioritized over
damage modes that are unlikely to occur or are of little
consequence in terms of safety. An RBI procedure is developed based on the assessment of typical damage modes
for the bridges being assessed that specifies the maximum
inspection interval.
Inspections are conducted according to the RBI procedure
developed through this process. The RBI procedure differs

Figure 1. Schematic diagram of the


RBI process.

141

tial outcomes or consequences of damage are also assessed.


Based on the assessment of the OFs and CFs for the various elements of the bridge, an inspection procedure is established, including the interval and scope for the inspection.
Criteria for reassessment of the inspection procedure are also
developed based on the assessment. The criteria for reassessment are typically based on conditions that may change as
a result of deterioration or damage, and that may affect the
OFs for the bridge. The RBI practice is then implemented in
the subsequent inspection of the bridge. Inspection results
are assessed to determine if any established criteria have been
violated, or if conditions have changed that may require a
reassessment of the OFs. If such changes exist, a reassessment
of the OFs is completed and the inspection practice modified
accordingly.
The method of determining the inspection interval, or time
period between inspections, is shown schematically in Figure2. The interval is based on the RAP assessment of the OFs
and the CFs, plotted on a simple two-dimensional reliability
matrix as shown in the figure. The OFs and CFs are used to
place typical damage modes in an appropriate location on the
matrix. In this figure, the horizontal axis represents the CF as
determined for a particular damage mode for a given bridge
element. The vertical axis represents the outcome of the OF
assessment for a given damage mode for the given element.
Damage modes that tend toward the upper right corner of the
matrix, meaning they are likely to occur and have high consequences if they did occur, require shorter inspection intervals
and possibly more intense or focused inspections. Damage
modes that tend toward the lower left corner, meaning they
are unlikely to occur, and/or consequences are low if they did
occur, require less frequent inspection. This is simply a rational approach to focusing inspection efforts; inspections are
most beneficial when damage is likely to occur and important
to the safety of the bridge; inspections are less beneficial for
things that are very unlikely to occur, or are not important to
the safety or serviceability of the bridge.

Figure 2. Reliability
matrix for determining
maximum inspection
intervals for bridges.

Through this process, individual bridges, or families of


bridges of similar design characteristics, can be assessed to
evaluate inspection needs from a reliability-based engineering
assessment of the likelihood of serious damage occurring, and
the effect of that damage on the safety of the bridge. The methodology can be applied throughout a bridge inventory, or to
portions of a bridge inventory. Suitable Quality Control (QC)
and Quality Assurance (QA) procedures should be utilized to
ensure consistency.
The RBI approach considers the structure type, age, condition, and operational environment in a systematic manner
to provide a rational assessment process for inspection planning. A documented rationale for the inspection strategy utilized for a given bridge is developed. The damage modes most
important to ensuring the safety of the bridge are identified
such that inspection efforts can be focused to improve the
reliability of the inspection results.
The sections that follow describe the key elements of the
RBI practices for bridge inspection. Section 3.3 provides
background data underlying the RBI process, including reliability concepts such as POF, the reliability theory applied
with the RBI process, damage modes and deterioration
mechanisms considered in the analysis, and typical lifetime
behavior characteristics that support the RBI approach. This
section also highlights the differences between the reliability
theory applied for inspection planning, and those traditionally applied for structural design codes. Section 3.4 discusses
key elements of the Guideline developed under the research
and initial testing of some of the processes developed. Finally,
Section 3.5 describes data needs and resources to support the
RBI analysis.

3.3Reliability
A key element in the RBI process is to understand the
meaning and role of reliability in the context of determining
inspection needs and inspection planning. This section of the
report provides supporting data and background information regarding important aspects of reliability and its underlying theories, and how these support RBI.
Reliability is defined as the ability of an item to operate safely
under designated operating conditions for a designated period
of time or number of cycles. The inspection practices documented in the Guideline are based on the concepts and theories of reliability. The reliability of a bridge element is defined
in terms of its safe operation and adequate condition to support the serviceability requirements for bridges. This definition is broader and more applicable to determining bridge
inspection needs than structural reliability estimates, which
are typically defined as a function of the load-carrying capacity of the structure and notional POF estimates. The challenge

142

with applying theoretical structural reliability concepts, such


as those used in modern design specifications, is that the
envisioned damage mode (loss of load-carrying capacity)
represents only a portion of the required information needed
from a bridge inspection. From the perspective of practical
bridge inspection, safe operation includes strength considerations, but also includes a variety of serviceability limit
states that may be related in some way to strength considerations, but are not direct measures of strength. Serviceability
considerations such as local damage that can affect traffic,
deflections and cracking, and loss of durability characteristics need to be assessed through periodic inspections, even
if the effect on structural capacity, and therefore structural
reliability, is nominal. Additionally, existing required load
ratings provide structural analysis in terms of load capacity for bridges (4). These ratings generally provide limited
insight into theinspection needs for a bridge, although the
engineering analysis considers certain inspection results,
such as section loss, in the analysis.
Several methods and processes have been suggested for the
assessment of in-service bridge reliability and the estimation
of inspection requirements based on structural reliability,
and these were studied during the course of the research.
Research based on structural reliability theory for the development of inspection strategies, repair optimization, and
updating bridge reliability estimates based on visual inspections has been performed (58). Significant work in the area
of applying structural reliability theory to highway bridges
was reviewed during the course of the research, and detailed
review is included in the project interim report (3, 7, 915).
The conclusion reached based on the review of this literature
was that these approaches were not currently implementable
for highway bridge inspection, due to several factors. First,
structural reliability models and probabilistic analysis does
not typically capture the serviceability limit states critical to
identifying in-service bridge inspection needs. Second, structural reliability models are highly theoretical in nature, and
the complexity of analysis required for even a simple structure
makes application to the diversified bridge inventory in the
United States impractical. Finally, the results of the structural
reliability assessments are often based on POF estimates that
are notional and design-based, such that significant uncertainty
would result from mapping these results to inspection needs
for specific bridges.
However, the underlying concepts of reliability could be
applied for the purpose of bridge inspection if appropriate
and implementable methodologies for estimating reliability
of bridges or bridge elements were developed. These methodologies need to consider the serviceability requirements for
bridges and bridge inspection, and define reliability appropri-

ately such that it can be assessed based on inspection results


and anticipated future deterioration. This analysis could then
be applied as one component of an inspection planning process that includes an assessment of the consequences associated with failure due to specific damage modes (1619).
Based on the analysis of the research on reliability methods,
the research team pursued a path to develop a semi-quantitative,
reliability-based framework for inspection practices. The key
elements of developing that methodology included identifying the reliability theories to be implemented to evaluate
bridges, and an appropriate description of failure to assess
when a bridge element is no longer performing adequately,
and hence has reduced reliability. The following sections
describe briefly the underlying reliability theory utilized
in the RBI Guideline, and the definition of failure used.
Damage modes and deterioration mechanisms that cause a
bridge element to deteriorate into the defined failure state
are discussed, and the overall concept of matching inspection needs to bridges during different stages of typical
in-service behavior are described.

3.3.1 Reliability Theory


Reliability is defined as the ability of an item to operate
safely under designated operating conditions for a designated
period of time or number of cycles. For bridges and bridge elements, reliability typically decreases as a function of time due
to deterioration and damage accumulated during the service
life of a bridge, for example, corrosion of steel elements in a
bridge that develops over the service life of the bridge, resulting in increasing damage over that service period. The likelihood of failure typically increases with time such that the
reliability of a bridge or bridge element can be expressed as:
R ( t ) = Pr (T t )
Where R(t) is the reliability, T is the time to failure for the
item, and t is the designated period of time for the items
operation. In other words, the reliability is the probability(Pr)
or likelihood that the failure time exceeds the operation time.
Sometimes, the likelihood is expressed as a probability density function (pdf) that expresses the time to failure of an
item (T) as some generic distribution, such as normal, log
normal, etc. (13, 15, 20). This distribution can be used to calculate a POF function, F(t), to express the probability that the
item will fail sometime up to time t. This time-varying function describes likelihood of failure up to some given time, or
the unreliability of the item, and the reliability is then:
R (t ) = 1 F (t )

143

In other words, the reliability is the probability that the


item will not fail during the time period of interest. The challenge for RBI was to determine an appropriate and practical
method of estimating the probability, or likelihood, of failure
described by the function F(t). This requires a definition of
what is meant by failure for a bridge element or structure.
It also requires an appropriate time interval over which an
effective and meaningful assessment can be accomplished
given the diversity in materials, designs, and operational
environments included across the bridge inventory.
When a large population of test data for identical or near
identical components exposed to the same operational environment are available, a probability function describing
the failure characteristics of the component may be determined and verified based on test results. This can provide
a quantitative frequency-based estimate of the POF that
indicates the number of events (failures) expected during
a given time period. However, such test data are generally
unavailable for bridges, because design, construction quality, and operational environments vary widely, and failures
are rare. A suitable probability distribution may be assumed
when test data are not available, but verifying the accuracy of
such a distribution can be difficult for complex systems like
highway bridges, where design and construction methods
are constantly evolving, operational environments vary, and
performance characteristics are also evolving. As a result,
past performance of similar elements of a bridge may not
be indicative of future performance, and the applicability
of an assumed function to a specific bridge is unverifiable,
since the lifetime failure characteristics described by the
assumed function describe events that have not yet occurred.
If designs, construction practices, and materials were not
evolving over time, this might be more practical, but this is
not the case for highway bridges.
Under conditions for which data to adequately characterize anticipated future behavior is limited, or where failure is
rare, engineering judgment and experience can be used to
estimate the expected reliability of a specific bridge within
a given operational environment (2123). Under these circumstances, the POF is determined based on qualitative or
semi-quantitative analysis and the probability is based on
degree of belief, rather than frequency. To make such decisions, individuals with expertise and experience with typical
performance characteristics, under a specific set of operational environments, is required. Utilizing expert judgment
and expert elicitation is a common method of characterizing
the reliability of components or systems for the purpose of
assessing inspection needs (2124). Such engineering judgment and knowledge provides data when quantitative data
are missing, incomplete, or inadequate. In the RBI method-

ology, expert elicitation is used as a process for estimating


the anticipated likelihood of failure for bridge elements, and
hence their reliability, over a given time period of 72 months.
The following sections describe the definition of failure,
damage modes and deterioration mechanisms, and typical
lifetime performance characteristics that are underlying the
RBI process analysis.

3.3.2Failure
A key step in assessing the reliability of a bridge element is
understanding how and why elements fail, and the typical
deterioration mechanisms that cause the elements to fail.
The damage modes and deterioration mechanisms that typically affect bridge elements are well known, in most cases.
For example, corrosion is obviously a significant deterioration mechanism in concrete and steel bridge elements that
causes them to fail. The likelihood of the failure occurring
in some future time interval depends on attributes of the element, such as its materials of construction, design, durability,
and current condition, as well as what conditions are used
to describe an element as failed. For bridges, catastrophic
collapse would be one obvious condition that could be used to
define failure, but such failures are very rare. Important concerns for bridge inspections extend well beyond simply avoiding rare catastrophic failures. Ensuring the safety of the bridge,
in terms of structural capacity, serviceability, and safety of
the traveling public are important factors in determining the
inspection needs of a bridge.
Therefore, failure requires a suitable definition that captures the need to ensure the structural safety of the bridge,
the safety of travelers on or below the bridge, and the serviceability of the bridge. Failure, utilized in this context, is defined
as when an element is no longer performing its intended function to safely and reliably carry normal loads and maintain
serviceability. For example, a bridge deck with severe spalling
may represent a failed condition for the bridge deck even
though the deck may have adequate load-carrying capacity, because the ability of the deck to reliably carry traffic is
compromised. Therefore, for the case of reliability assessments
for determining bridge inspection needs, it was necessary to
adopt a commonly understood definition of failure that considers common deterioration patterns in bridges and that can
effectively be assessed through the inspection process. Additionally, failure must be defined in a commonly understood
manner that can be readily assessed, is consistent with the historical experiences of bridge managers, and is sufficiently
general to be easily applied across the broad spectrum of
design characteristics and elements that exists across the
bridge inventory. To meet this need, the NBIS condition

144

rating of 3, serious condition, was chosen as a general, durable,


and readily understood definition of failure. Bridge elements
that have deteriorated to this extent may no longer be performing their intended function, and remedial actions are typically
planned to address such conditions. It is not envisioned that
any bridges or bridge elements assessed using a reliabilitybased approach are allowed to deteriorate to this condition.
Rather, inspection intervals are adjusted to ensure that the
likelihood of failure in the time intervals between inspections
always remains low.
The subjective condition rating of 3 is defined within the
Recording and Coding Guide (25) as follows:
NBIS Condition Rating 3: SERIOUS CONDITION: Loss of section, deterioration, spalling or scour have seriously affected primary structure components. Local Failures are possible. Fatigue
cracks in steel or shear cracks in concrete may be present.

In terms of the AASHTO Bridge Element Inspection Guide,


this condition generally aligns with elements in Condition
State (CS) 4, serious (26).
These condition descriptions are widely understood and
there is significant past experience in the conditions warranting
a rating of 3 throughout the bridge inventory for the myriad of
different materials and design characteristics that exist. This
condition description provides a practical frame of reference
for assessing the likelihood of failure in some future time
period. For example, one could readily assess if a bridge deck
that currently has a condition rating of 7, and has durable
attributes such as adequate concrete cover and epoxy-coated
reinforcing steel, was very likely, or very unlikely to deteriorate
to a condition rating of 3 in the next 72 months. If the deck is
very unlikely to deteriorate to a failed state during that time
interval, repeated inspections of the deck may yield little or
no benefit. On the other hand, if the deck were in poor condition, with a condition rating of 4, it may be more likely
to fail during this time period, and more frequent inspections are necessary to monitor the deterioration and identify
repair needs.

3.3.3Damage Modes and Deterioration


Mechanisms
The failure state described above is typically reached as the
result of the accumulation of one or more forms of damage.
For example, a deck may reach the failed state because of
widespread spalling; a steel beam may reach that state as the
result of severe section loss. These typical forms of deterioration in bridges are observable in a visual assessment of the
bridge, or sometimes with the assistance of a nondestructive evaluation technology (NDE). The observable effects on
which a condition assessment is normally based are forms

of damage, or damage modes. Damage modes are typically


assessable through the inspection process and their extent or
degree recorded in the inspection results. Spalling, cracking,
scaling, sagging, etc. are damage modes.
Damage modes are normally the result or manifestation
of a deterioration mechanism, such as corrosion or fatigue.
Deterioration mechanisms describe the path to failure, and
may occur at different rates depending on factors such as
operational environment and loading patterns. For example,
a concrete bridge deck may fail due to the damage mode of
concrete spalling, and the deterioration mechanism is corrosion. If the deck is located in an aggressive environment, the
corrosion mechanism may be fast acting, if in a benign environment, the mechanism may be slow acting. Similarly, if the
damage mode is cracking in a steel element, and the cracking
results from the deterioration mechanism of fatigue, then the
rate at which the damage mode will progress depends on the
cyclic loading of the bridge. If the bridge has very low average
daily truck traffic (ADTT), then the likelihood of the damage mode progressing is lower than if the ADTT were high.
However, if the damage mode is cracking and the deterioration mechanism is constraint-induced fracture (CIF), the
progression of the damage mode may only depend on the
susceptibility of the weld detail to CIF.
Within the RBI process, it is important to separate the
damage mode from the deterioration mechanisms such that
suitable attributes or characteristics can be appropriately
identified. For example, if the damage mode is spalling in a
bridge deck, the deterioration mechanism could be corrosion
of embedded reinforcing steel, or could be debonding of an
overlay. Obviously, the attributes affecting how likely it is that
debonding will occur differ from those that affect how likely
it is that corrosion damage may occur, even though the resulting damage may have very similar effects on the serviceability
of the deck.

3.3.4 Lifetime Performance Characteristics


Part of the overall assessment of the reliability of a bridge
element is an understanding of the typical lifetime behavior
of engineering components. Generally, failure patterns can
be described by a bathtub curve such as that shown in
Figure 3, which represents the failure rate, or POF, as a function of the time. The bathtub curve shows the initial failure
of new components due to defect (infant mortality), the useful life period, and the wear-out period. For bridges, the infant
mortality portion of the bathtub curve illustrates the effects
of construction errors or flaws, which typically become evident in the early life of a bridge. One of the purposes of QCs
and inspections during the construction phase of a bridge
is to reduce the infant mortality rate, that is, to ensure there
are not defects in the structure from construction errors that

145

Figure 3. Plot of the bathtub probability


curve.

will lead to a shorter than expected service life. Following the


period in which infant mortality may occur, elements typically have long service lives and failures are rare. Toward the
end of the service life, when elements are in advanced stages
of deterioration, the likelihood of failure can increase substantially. As a result, more frequent and thorough inspections may be necessary to monitor deterioration and identify
repair needs. The bathtub curve shows schematically the
typical performance of engineered components: the shape
and timeline of the curve for specific bridge elements obviously
depends on the attributes of the element, including the design
characteristics, typical construction quality, operational environment, management and maintenance practices, etc. Among
the purposes of RBI or any other life-cycle management
system that includes inspection is the reduction of the wearout rate by finding and repairing or replacing components
before they fail, reducing unnecessary or unjustified inspec-

tion efforts, and optimizing the utilization of inspection


resources. Inspection needs are typically lower during the
useful life of elements, when failures are rare, and increase as
the failure rate increases as the result of deterioration mechanisms that manifest in damage.
Many different methods are available to model failure processes and determine failure rate characteristics such as those
shown in Figure 3, from qualitative to quantitative including
hybrid methods. Qualitative methods would include expert
judgment; hybrid methods would include methods like
Markov Chain models, which use expert opinions and empirical data to estimate transition probabilities (27). Quantitative methods can range from fully empirical (using statistical
fits to test or field data) to fully physics-based (using physical models of failure processes). Weibull and log-normal
statistics have both been used to describe failure processes
that are driven by forces such as fatigue, wear, and/or corrosion. Given a sufficiently large population of engineered
structures and the same driving forces, their rate of failure
(or equivalently the POF at any time) can often be described
by Weibull or log-normal statistics. Thus, if items are cheap
and easy to test, a statistical description of their failures can
be created and used to predict the behavior of similar items
in the future. However, for bridges, characteristics of the elements and their environment vary widely and are difficult to
capture within such models, particularly when considering
the needs of a specific bridge. For example, Table 1 shows
variables used for probabilistic modeling of bridge reliability from some common literature resources and indicates
the level of data resources that need to be either determined
empirically, or estimated using statistical tools and probability functions (9, 10, 14, 28). As this table indicates, the magnitude of data that needs to be either collected or assumed is
significant. The assumptions required to effectively estimate
such a large number of properties and characteristics
requireverification, and may vary widely across different

Table 1. Variables used for probabilistic estimates of time-varying


reliability.
Concrete cover

Corrosion rate

Time to corrosion initiation

Workmanship

Crack width

Prestress steel strength and


modulus

Critical crack width

Prestress losses

Crack depth

Impact factor

Concrete strength and


modulus
Reinforcing steel
strength and modulus
Shrinkage of concrete

Cracking density

Thickness

Loading rate
Surface chloride
concentration
Critical chloride
concentration

Dead load
Truck live load
Water-cement ratio

Chloride diffusion

Area of reinforcing steel in


concrete
Flexural forces
Shear forces
Load distribution factors
Reinforcement spacing

146

bridge design, operational environments, and construction


practices. Verification of the assumption requires observation of bridge performance over its service life, and therefore,
by definition, cannot be determined in time to be usefully
applied. It is also notable that among the many parameters
assembled to estimate time-dependent reliability, bridge joint
condition is not among them. However, practically, this factor alone may outweigh all of the others in terms of assessing
the expected deterioration patterns and rate for a bridge.
Additionally, because design and construction processes
are evolving, elements that have the same role in different
bridges often do not share key design features or operational
environments that could affect their long-term performance.
This makes estimating the many factors shown in Table 1
even more challenging and impractical across an inventory
that includes 600,000 bridges and a multitude of operational
environments. Therefore, expert judgment is required to
consider the role and significance of specific design and environmental features for specific bridges, and to estimate future
performance effectively.

3.4 Key Elements of RBI


This section of the report provides an overview of key elements of the RBI process described and detailed in the Proposed Guideline for Reliability-Based Inspection Practices
developed through the research. This includes a description
of the OF, the CF, inspection procedures for RBI, and the RAP.

3.4.1 The OF
Within the RBI process, an estimate of the POF for a given
bridge element is expressed as an OF. This factor is an estimate of the likelihood of severe damage occurring in a specified time interval, considering the likely damage modes and
deterioration mechanism acting on the element. Key attributes of the element that affect the likelihood are considered and documented to support the estimate. This section
describes the approach and methodology for estimating the
probability, or likelihood, of failure for bridge elements for
the purpose of inspection planning.
There are a variety of methodologies for estimating the
expected performance of components or elements. These
range from fully quantitative methods to fully qualitative
methods. For example, the American Petroleum Institutes
Recommended Practice 581 has, for certain critical components, empirical equations that estimate the POF for the
component given certain attributes of the component and
its operational environment (29). These empirical equations
include factors associated with the attributes of specific components and are used to calculate the expected POF over some
defined time period. In other cases, physics-based models for

damage such as fatigue cracking are combined with industrial


modeling tools to estimate the POF for specific components
or systems (17, 3033). For cases in which historical data may
be scarce, where systems are complex and/or evolving such
that relevant historical data are unavailable, expert judgment
and expert elicitations can be used (21).
To develop an estimate of the POF over a certain period,
several factors need to be considered, including what constitutes a practical definition of failure, as described above,
over what time period the assessment can be made, and what
resolution is required for the estimate. Often, estimates utilized in reliability analysis are simply order of magnitude
estimates, or even ranges, over which the POF is expected to
fall. For example, ASME guidelines suggest first-level qualitative analysis can be achieved using a simple three-level scale
shown in Table 2 (21). An estimate of the annual POF associated with a qualitative ranking is also provided. In this context
(for this industry) a high POF is intended to represent failure
rates on the order of a 0.01 or 1 in 100, while moderate probability (likelihood) is intended to cover 2 orders of magnitude
from 0.01 to 0.0001, with low probably being less than 0.0001.
In moving from totally qualitative to semi-quantitative analysis,
the order of magnitude of the failure rate may be estimated,
and these numerical values provide a mapping of qualitative to
quantitative rankings. In practical applications, even if quantitative methods are used, the estimated POFs are typically considered to be, at best, order of magnitude estimates, due to the
inherent variation and uncertainty in engineered systems.
For application for the RBI assessment for highway bridges,
existing industrial approaches were considered as a basis for
developing appropriate methodologies for estimating reliability for highway bridges elements. This required that an appropriate time interval be determined over which an assessment
for the POF could be made, based on available data and engineering factors. Appropriate categorizations or qualitative
scales to effectively describe that reliability were developed
for use as part of a reliability-based assessment.
3.4.1.1 Assessment Interval
Given the typically long service life of a bridge and the slow
rate of deterioration mechanism such as corrosion, annual
POF estimates such as those described above may have little
Table 2. ASME POF rankings using a
three level scale.
Possible Qualitative Rank
Low
Moderate
High

Annual Failure
Probability
<0.0001 (1/10,000)
0.0001-0.01 (1/10,000
1/100)
>0.01 (1/100)

147

meaning and vary widely according to the assumption made


in a particular analysis. Additionally, failure is not typically
well defined as is the case, for example, with a pipe or valve.
If the pipe leaks, it is failed, if the valve fails to open when
required, it is failed. But with elements in a bridge, the majority of deterioration mechanisms extend over long timeperiods, fracture being an exception, and the failure state itself
can be subjective (34). Elements may reach a state that meets
the definition of failure and stay in that condition for
some number of years. Therefore, it is more appropriate to
describe how likely it is for deterioration or damage to occur
to the extent that an element deteriorates into a serious or
failed condition. For the RBI process for bridge inspection,
an OF is used to represent a qualitative measure of this likelihood over a time interval of 72 months. This time period
was determined based on engineering factors that included
prior research, experience, expert judgment, and data from
corrosion, damage and deterioration models (13, 3543).
For example, commonly available corrosion models indicate
that significant periods of time transpire between construction of a bridge and initiation of corrosion, particularly in
environments that are not aggressive (i.e., little or no use of
de-icing chemicals and no marine exposure). Once initiated,
corrosion may take a significant period of time to manifest in
damage, depending on factors such as bar spacing, cover, concrete material properties, and environment. Estimates for damage progression typically range from 6 years on the low end,
for uncoated rebar in typical concrete structures, to 20 years
or more for epoxy-coated bars. For steel elements, although
corrosion damage can be severe, the rate at which corrosion
damage occurs is actually very slow, typically less than 0.004
0.006in./year, even in moderately aggressive environments
(37, 44, 45). Therefore, the amount of section loss that could
occur during a 72-month interval is nominally less than
1
16 of an inch, assuming two sides of a steel plate were corroding equally, at a relatively high rate of 0.005 in./yr. Section loss
on this order of magnitude would not be considered serious.
Therefore, it is practical to assess the likelihood of damage progression occurring over a time frame of 72 months, because
the likelihood is low that these deterioration mechanism could
result in a bridge element deteriorating from a good condition to a serious condition during such a time interval. It
should be noted that this interval of 72months is an assessment
interval over which the reliability of an element is estimated for
the purpose of assessing inspection needs. Inspection intervals
may be significantly less than 72months when existing damage
is present, or the attributes of an element suggest the likelihood
of damage developing is high.
The time period of 72 months is also considered a time
period for which an engineer could reasonably estimate future
performance within four fairly broad categories, ranging from
remote to high, based on key attributes that describe the

design, loading, and condition of a bridge or bridge element.


The interval provides a suitable balance between shorter
intervals, when the POF could be unrealistically low do to
the typically slow progression of damage in bridges, or longer
intervals, where uncertainty would be increasingly high. For
example, if an engineer was asked to predict if a deck currently
rated in good condition was likely to progress to a serious state
in 1 year, that estimate would be very low, since deterioration
mechanisms are slow acting. However, if the time period of
10 years were used, the uncertainty could be very high. The
time interval was selected in part to provide a suitable balance
over which damage progression could be reasonably predicted
based on engineering assessment and rationale.
3.4.1.2 OF Categorization
The OF is a qualitative ranking or categorization of the
likelihood that an element will fail during a specified time
interval. A four category, qualitative scale was developed for
estimating the OF for RBI practices. The scale ranges from
remote, when the likelihood is extremely small such that it
would be unreasonable to expect failure, to high, where the
likelihood of the event is increased. The categories and associated verbal descriptions are shown in Table 3.
The OF is determined by expert judgment considering key
characteristics, or attributes, of bridge elements. Attributes
are characteristics of a bridge element that contribute to the
elements reliability, durability, or performance. These attributes are typically well-known parameters affecting the performance of a bridge element during its service life. These includes
relevant design, loading, and condition characteristics that are
known or expected to affect the durability and reliability of the
element. These attributes are identified and assessed through
the expert elicitation process.
Numerical ranges that could be used to describe the OF scale
quantitatively are shown in Table 4. Such numerical values
provide ranges or target values for the qualitative rankings
that could be used to map quantitative data, if it is available,
to the qualitative rating scales. Failure of a bridge element is
a relatively rare event, and design and construction details
vary widely. As a result, relevant and verifiable frequencybased probability data are scarce, as previously discussed.
The numerical values shown in Table 4 are target values that
Table 3. OF rating scale for RBI.
Level
1

Category
Remote

2
3
4

Low
Moderate
High

Description
Remote likelihood of occurrence,
unreasonable to expect failure to occur
Low likelihood of occurrence
Moderate likelihood of occurrence
High likelihood of occurrence

148
Table 4. OF categories and associated interval estimates of POF.
Level

Qualitative
Rating

Description

Likelihood

Expressed as a
Percentage

Remote

Remote probability of
occurrence, unreasonable to
expect failure to occur

1/10,000

0.01% or less

Low

Low likelihood of occurrence

Medium

High

Moderate likelihood of
occurrence
High likelihood of occurrence

can be used to map such data or models to the qualitative


scales used in the analysis. For example, data from PONTIS
deterioration curves or from probabilistic analysis, or other
deterioration models, could be incorporated directly into
the assessment of the OF using these scales. These numerical
categories can also provide a framework for future development of models or data derived from analysis of the deterioration patterns in a particular bridge inventory.
The quantitative description can be also be used as a vehicle for expert elicitation by using common language equivalents for engineering estimates. For example, if you asked an
expert to estimate the probability of serious corrosion damage (widespread spalling, for example) for a particular bridge
deck given its current condition, a common engineering
response might include a percentage estimate, for example,
less than 0.1% chance or less than 1 in a thousand. This estimate can then be mapped to the qualitative scale as being
low. Such estimates are typically very conservative, particularly for lower, less likely events.
3.4.1.3 Method of Assessing OFs
OFs are determined through expert elicitation by the RAP
assembled by a bridge owner. The RAP provides experience
and knowledge of the performance of materials, designs, and
construction quality and methods within a specific operational environment. This knowledge and experience is used
to categorize the OF considering credible damage modes and
deterioration mechanisms for bridge elements.
The assessment is conducted by identifying critical design,
loading, and condition characteristics, or attributes, that
affect the reliability and durability of the elements. For example, consider the damage mode of spalling due to corrosion
damage in a concrete bridge deck. A bridge deck may have
good attributes, such as being in very good condition, having adequate concrete cover, epoxy-coated steel reinforcing,
and minimal application of de-icing chemicals. Given these
attributes of the deck, it may be very unlikely that severe damage (i.e., failure) would occur in the next 72 months. This

1/10001/10,000
1/1001/1,000
>1/100

0.1% or less
1% or less
> 1%

is based on the rationale that the deck is presently in good


condition, and has attributes that are well known to provide
resistance to corrosion damage. As such, an OF of Low or
Remote might be used to describe the likelihood of failure
due to this damage mode. Alternatively, suppose the deck is
in an environment where de-icing chemicals are frequently
used, the reinforcement is uncoated, and the current rating
for the deck is a 5, Fair Condition, indicating that there are
signs of distress in the deck. Based on this rationale, the likelihood of serious damage developing would be much greater,
resulting in an OF rating of Moderate or High. Past experience with decks of a similar design, combined with engineering judgment, can be used to support the assessment of
the specific OF for a given deck.
These attributes can be generally grouped into three categories: Design, Loading, and Condition attributes. Design attributes of a bridge element are those characteristics of the element
that describe its design. Design attributes are frequently intrinsic characteristics of the element that do not change over time,
such as the amount of concrete cover or material of construction. In some cases, preservation or regular maintenance activities that contribute to the durability of the bridge element
may be a design attribute, such as the use of penetrating sealers as a preservation strategy.
Loading attributes are characteristics that describe the loads
applied to the bridge element. This may include structural
loading, traffic loading, or environmental loading. Environmental loading may be described in macro terms, such as the
general environment in which the bridge is located, or on a
local basis, such as the rate of de-icing chemical application on
a bridge deck. Loading attributes describe key loading characteristics that contribute to the damage modes and deterioration mechanisms under consideration.
Condition attributes describe the relevant bridge element
conditions that are indicative of its future reliability. These
can include the current element or component-level rating,
or a specific condition that will affect the durability of the element. For example, if the damage mode under consideration
is concrete damage at the bearing, the condition of the bridge

149

joint may be a key attribute in determining the likelihood that


corrosion will occur in the bearing area.
3.4.1.4 Screening Attributes
Attributes can also be identified as screening criteria that
identify certain characteristics that have a predominate effect
on the reliability of an element. Attributes used for screening may be design, loading, or condition attributes. Screening
attributes are used to quickly identify bridges that should not
be included in a particular analysis, either because they already
have significant damage or they have attributes that are outside
the scope of the analysis being developed. Screening attributes
are typically attributes that:
Make the likelihood of serious damage occurring very high,
Make the likelihood of serious damage occurring unusually

uncertain, and
Identify a bridge with different anticipated deterioration

patterns than other bridges in a group.


The RAP must identify the appropriate value/condition for
the attribute to use as a screening tool. For example, if considering the likelihood that the steel bridge will suffer corrosion
damage that reduces its rating to a 3, and the current rating
is 4, the RAP may consider that such condition indicates that
there is an unusually high likelihood of further damage developing over the next 72-month period, and as such, use the
condition rating of 4 as a screen. In such a case, the analysis
can move forward to an assessment of the consequences of
the damage without further evaluation of the attributes that
contribute to the likelihood of damage, based on the screening item. Another example would be to screen steel beam
elements in bridges that have open decking. Since the open
decking allows drainage directly onto the steel beams, the
deterioration of these bridges would not be similar to steel
beams with typical concrete decks; these bridges are screened
from the analysis of steel beam bridges, as they may require
separate analysis. It may be appropriate to treat these bridges
as a separate group, developing the analysis to consider key
attributes of those bridges with open decking.
3.4.1.5 Ranking Attributes
Key attributes for a bridge element are identified by the
RAP and used to assess the appropriate OF for the given element and damage mode being considered. This assessment
is supported through an empirical scoring procedure that
provides a rational method of estimating the OF category.
The attributes identified are ranked according to the importance of each attribute in assessing the reliability of a certain
bridge element. For example, for attributes that play a

primary role in determining the likelihood of damage, a


scale of 20 points could be used, 15 points for an attribute
that has a moderate role, and 10 points for an attribute that
plays a minor role. For the damage mode of corrosion in a
steel beam, for example, a leaking joint which results in drainage of de-icing chemicals directly onto the superstructure is
highly important in assessing the likelihood of serious corrosion damage occurring. Therefore, this attribute may be
assigned a 20 point scale by the RAP. The RAP may consider
age of the structure to contribute to the likelihood of corrosion damage, but to a much lesser extent relative to a leaking
joint, and assign a 10point scale.
Once the overall importance of the attribute is identified,
different conditions or situations may be described to distribute points appropriately based on the engineering judgment
of the RAP. Again using the joint as an illustration, if the joint
is leaking or can reasonably be expected to be leaking during
the time interval, it will have the highest effect and be scored
the full 20 points. If the joint is debris-filled or exhibiting
moderate leakage, a score of 15 points may be appropriate, if
there is a joint, but it is not leaking, a score of 5 points may be
assigned. If the subject bridge is jointless, a score of 0 points
may be used. The exact scoring for a given attribute may vary
according to the design characteristics or operational environment of a particular bridge inventory. The key attributes
and ranking scores are then used to develop a simple scoring
process that ranks the reliability characteristics of a particular
element, for a given damage mode, as a rational means of
assessing the appropriate OF.
The scoring methodology is intentionally flexible to adjust
to the needs and requirements of different bridge inventories, while still providing a systematic process to document
rationale for the OF assessment. It is not a one size fits
all approach, but rather intended to conform to the varying needs of different operational environments and bridge
inventory characteristics. The commentary section of the
Guideline, Appendix E, provides suggested scoring and rationale for more than 50 common attributes that might be
identified by a RAP assessment of concrete and steel bridges.
Alternatively, the RAP may identify additional attributes that
meet the needs of a particular inventory, and develop rationale explaining the purpose and assessment process for the
attribute. Suggested scoring weights for the attributes may
also vary according to the needs and experiences within particular operational environments. Calibration of the scoring
process is obviously required to ensure the overall assessment
of attributes is consistent with engineering judgment.
Certain key attributes should be identified as part of criteria
for reassessment of bridge inspection requirements, following
subsequent inspections. These attributes are typically associated with condition, which may change over the service life
of the bridge as deterioration occurs. When changes in these

150

condition attributes can result in a change in the likelihood of


a given damage mode resulting in failure (i.e., the OF), reassessment of the inspection requirements is necessary.
Several illustrative examples of attribute scoring are also
provided as guidance in making the assessments. This includes
scoring sheets for tabulating scores for different elements of
bridges, and using those scores to determine the OF. However, once attributes and attribute rankings for bridge elements
are determined by an RAP, the scoring may be more readily
accomplished by integrating or developing software for scoring
characteristics of bridge elements more efficiently. Some of the
attributes identified by the RAP may already be stored in existing databases and bridge management systems; others may
need to be acquired from inspection reports, bridge plans,
and other sources.
3.4.1.6 Use of Surrogate Data
For many bridges, the use of surrogates for the attributes identified in the reliability analysis may be considered
to improve the efficiency of the analysis for larger families
of bridges. As used herein, surrogate refers to specific data
that can be used to either infer or determine another piece of
information that is required for the reliability assessment. For
example, assume a fracture critical bridge was designed and
built in the year 2000, which is well after the implementation
of the AASHTO/AWS Fracture Control Plan. This information can be used to determine that the steel must at least meet
certain minimum toughness requirements, and the bridge
meets modern fatigue design requirements. Note that this
was determined only from the date of construction and with
no detailed review of the design calculations or specifications.
As stated, the use of surrogates is particularly attractive
when identifying and assessing a family of bridges. Design and
loading attributes identified by the RAP are typically static in
nature, that is, they do not change over time. The condition
attributes will typically change over time, as damage accumulates and deterioration mechanisms manifest. However, when
elements are in generally good condition, specific condition
attributes identified by the RAP may not require individual
assessment for each bridge or family of bridges; the previous inspection results can simply be used as a surrogate for
the individual attributes. This will typically allow for larger
groups of bridges of similar design to be grouped into a particular inspection interval, based on the criteria developed by
the RAP. For example, again considering steel bridges built to
modern design standards, it is known that the design attributes that would increase the likelihood of fatigue cracking
and fracture have been mitigated through improvements in
the design, fabrication, and construction process. The condition attributes that are required to assess the reliability of
the element would include the presence of fatigue cracks due

to out-of-plane distortions, fatigue cracking due to primary


stresses, and corrosion damage. However, if the component
rating is 7, in good condition according the NBIS scale, or
CS 1 in an element-level scheme, the existing ratings can be
used as a surrogate for the condition attributes. Note: This
assumes the inspection result is from an RBI procedure, i.e., the
inspection was capable of identifying if fatigue cracks existed.
This allows all bridges that are of this same rating (and similar design and condition attributes) to be treated collectively
in a process that does not require much detailed analysis of
individual bridges. If the condition rating or condition state
changes, then the bridges can be reevaluated according to the
RAP criteria. If the condition does not change between periodic inspections, reassessment may not be necessary.
It is important to note that this process is significantly different than assigning an inspection interval based simply on
the current condition of the bridge; for example, deciding to
inspect all steel bridges with rating of 7 on a longer interval than
all of those rated a 6. The RAP analysis forms a rationale that
identifies not only the current condition attributes that affect
the reliability of the element, but also the design and loading
attributes of the bridge or bridge element that affect the potential for damage to occur. In other words this RAP evaluation
forms an engineering rationale for the decision-making process
that considers not only the condition of the element, but also
the damage modes and the potential for that damage to occur.
For element-level inspection schemes, the attributes identified by the RAP may map directly to an element and element
condition state. For example, consider that the RAP identifies
leaking joints as an attribute driving the likelihood of section
loss in the bearing area of a steel beam. The element condition state (joint leaking) is recorded in the inspection process
and can be used as a criterion for that attribute score. In some
cases, all of the attributes identified by the RAP as being critical to the likelihood of failure of an element may be included
in a comprehensive element-level inspection process; in other
cases, they may not.
For NBI-based inspection schemes, attributes identified by
the RAP may map to sub-element data collected in addition
to the required condition ratings for the primary components
of the bridge. This data could be used if it is collected under
a standardized scheme for rating and data collection for the
sub-elements. For the primary components, the generalized
nature of the component rating makes this more difficult for
specific attributes.

3.4.2CFs
The second factor to be assessed under the RBI process is
the CF, a categorization of the likely outcome presuming a
given damage mode were to result in failure of the element
being considered. The assessment of consequence is geared

151

toward assessing and differentiating elements in terms of the


consequences, assuming that failure of the element occurs.
It should be noted that failure of an element is not an anticipated event when using an RBI approach, rather the process of
assessing the consequences of a failure is merely a tool to rank
the importance of a given element relative to other elements
for the purpose of prioritizing inspection needs.
The CF is used to categorize the consequences of the failure
of an element into one of four categories, based on the anticipated or the expected outcome. Failure scenarios are considered based on the physical environment of the bridge, typical
or expected traffic patterns and loading, the structural characteristics of the bridge, and the materials involved. These
scenarios are assessed either qualitatively, through necessary
analysis and testing, or based on past experience with similar failure scenarios. The four-level scale used to assign the
CF is shown in Table 5. The CF ranges from low, used to
describe failure scenarios that are benign and very unlikely to
have a significant effect on safety and serviceability, through
catastrophic scenarios, where the threat to safety and life is
significant. Thus, both short-term (generally safety related)
and long-term (generally serviceability related) consequences
can be considered.
In assessing the consequences of a given damage mode
for a given element, the RAP must establish which outcome
characterized by the CFs in Table 5 is the most likely. In other
words, which scenario does he or she have the most confidence will result if the damage were to occur. Using the illustration of brittle fracture in a girder, it is obvious that the
most likely consequence scenario would (and should) be different for a 150 foot span two-girder bridge than for a 50 foot
span multi-girder bridge. For the short-span, multi-girder
bridge, an engineer may state with confidence that the most
likely consequence scenario is High and that Severe consequences are very remote for a multi-girder bridge, based on
his/her experience and the observed behavior of multi-girder
bridges. For the two-girder bridge, the consequence scenario
is likely to be Severe. As this example illustrates, the CF simply ranks the importance of the damage mode as being higher
for a two-girder bridge than for a multi-girder bridge. For
many scenarios, qualitative assessments based on engineering
judgment and documented experience are sufficient to assess

the appropriate CF for a given scenario; for others, analysis


may be necessary using suitable analytical models or other
methods. A series of more detailed criteria for specific elements
(i.e., decks, steel girders, P/S girders, etc.) are provided in the
Guideline that can be utilized during the assessment to determine the appropriate CF for a given element failure scenario.
These criteria, combined with owner-specific requirements
developed in the RAP or from other rational sources for assessing bridges and bridge redundancy, are then used to determine
the appropriate CF for a given scenario.

3.4.3 Inspection Procedures in RBI


Conducting a reliability-based assessment of the inspection needs for bridges requires specific information regarding
the current condition of bridge elements that allows for the
assessment of expected future performance. For example, to
determine the appropriate OF for corrosion damage in a steel
bridge element, one would have to know if corrosion damage
were currently present and to what degree or extent. Without
this information, it would not be possible to assess the likelihood of severe damage developing over the next 72 months.
Therefore, it is necessary under the RBI approach to perform
inspections that are capable of detecting and evaluating relevant damage modes in a bridge. The relevant damage modes
for specific bridge elements are identified through the RAP
analysis of the OF, and this assessment provides foundation
for the inspection scope and procedures to be used in the field
for future inspections. The thoroughness of the inspection
process is typically increased relative to, for example, component-level approaches that require only a single rating for a
component (superstructure, substructure or deck).
The methods or procedures used to conduct the inspection
must be capable of reliably assessing the current condition of
the bridge elements for the specific damage modes identified
through the RBI process. In many cases, visual inspection or
visual inspection supplemented with sounding may be adequate for conducting RBI. The inspections may be hands-on,
such that damage is effectively identified to support the reliability assessment. For example, when assessing the likelihood
of severe fatigue cracking in a bridge (the OF), it would be necessary to know if there were currently fatigue cracks. To make

Table 5. CFs for RBI.


Level

Category

Consequence
on Safety

Consequence on
Serviceability

Low

None

Minor

Moderate

Minor

Moderate

High

Moderate

Major

Severe

Major

Major

Summary Description
Minor effect on serviceability,
no effect on safety
Moderate effect on serviceability,
minor effect on safety
Major effect on serviceability,
moderate effect on safety
Structural collapse/loss of life

152

that assessment, sufficient access to the superstructure of a


bridge is required to determine if fatigue cracking is currently
present, obviously, and the inspection procedure must include
reporting the presence or absence of fatigue cracks. In some
cases, NDE techniques may be required within the inspection
procedure to allow for reliable detection of certain damage
modes identified through the RBI analysis. For example, if
the RAP identifies cracking in a bridge pin as a credible damage mode because a bridge has pin and hanger connections, a
visual inspection is inadequate. Because the surface of the pin
where cracking is likely to occur is not accessible, due to interference from the hanger plates, beam web and reinforcements,
ultrasonic testing (UT) or other suitable NDE technology is
necessary to allow for the cracking to be assessed.
The RAP analysis of the OFs and CFs provide a basis for the
inspection requirements to be used in the field, by identifying
credible damage modes and prioritizing these damage modes
based on their potential effect on safety and serviceability.
Based on the assessment of the OFs and the CFs, damage
modes for a bridge can be prioritized based on the product
of these factors:
IPN = OF CF
Where IPN = Inspection Priority Number. For example, if
the fatigue cracking has a moderate likelihood of occurring
and the consequence is severe, then the IPN would be 3 4
= 12. If fatigue cracking were moderately likely, but the consequence were only moderate (minor service disruption),
for example, if the bridge in question is a short-span, multigirder bridge with known redundancy, the IPN for that damage mode would only be 3 2 = 6. This process highlights the
damage modes that are most important, that is, most likely
to occur, and have the greater associated consequences if they
did occur. This information is included in the inspection procedure for the bridge, providing guidance to the inspectors
on emphasis areas for the inspection, based on the engineering analysis and rationale developed by the RAP.
It should be noted that the calculation of the IPN for
each damage mode identified in the process does not limit
the scope of the inspection to only those damage modes.
However, it provides a simple method of prioritization of
damage modes that are most important, based on a rational
assessment that incorporates bridge type, age, design details,
condition, etc., as well as the associated consequences. The
resulting outcome from the RAP analysis provides inspection requirements that are tailored to the specific needs of the
bridge and include a prioritization of the damage modes for
that bridge. This provides a more focused inspection practice that is based on an engineering assessment of the specific
bridge or bridge type in order to improve the effectiveness
and reliability of the inspection.

3.4.3.1 Reliability of Inspection Methods


For most RBI planning processes, such as those used for
assessing cracking in nuclear power plants or oil and gas facilities, the reliability of different inspection strategies or methods
is considered the assessment (21, 23, 29). For inspection technologies, reliability is typically defined by a measure of the ability of the technology to perform its intended function. Reliable
and effective inspection methodologies reduce the uncertainty
in the current condition of components, and therefore can
affect future POF estimates and rationale for a given inspection
interval. The reliability of specific inspection methods may be
quantified using probability of detection (POD) or other reliability analysis for a limited number of especially high-risk
components and damage scenarios. This may be justified based
on the significant risk associated with these facilities, including both the high cost and high environmental consequences
of certain failure modes. However, for more general assessments of risk, the effectiveness of inspections is qualitatively
described to rank various inspection approaches on a relative
scale using engineering judgment. For example, API has created a five-category rating system used for several components
described in API 581 (29). Inspection methods are qualitatively
categorized on a scale that ranges from A to E, with A being
highly effective and E being ineffective.
A similar approach was taken to develop guidance on the
reliability or effectiveness of inspection methods for typical
damage modes anticipated for common bridge elements.
Tables included in the Guideline indicate the reliability
of NDE technology for various damage modes for specific
bridge elements, such as steel beams, concrete decks, etc. The
reliability of the inspection method is described on a fourlevel qualitative scale and represented symbolically. Methods
that are generally unreliable for a given damage mode or
mechanism are described as Low and methods expected to
provide high reliability and effectiveness are High. The assessments of the reliability of inspection methods were made using
expert judgment, literature review, experience, and data from
other industries, where available (46). Information on the relative costs of different methods is also included as guidance. The
Technical Readiness Level (TRL) of different methodologies
is also provided and describes if the methodology is a commonly available tool that is readily accessible, if the method
is specialized such that specialized expertise is required for
implementation, or if the method is experimental in nature.
Presently, there is somewhat limited reliability data available
for many bridge inspection techniques and NDE technologies
applied for bridge inspection. In part this is because historically there has been little motivation to conduct such testing,
since the inspection intervals are uniform and generally do
not require any formal demonstration of effectiveness of the
inspection procedure. However, in an RBI approach, where

153

inspection intervals may be longer based on rational assessments of potential damage, inspection scopes may need to be
appropriately adjusted. As a result, determination of the reliability of the inspection method becomes a factor in the overall approach to the inspection process. Reliability data such as
that provided in the Guideline is expected to be refined and
developed over time, as the reliability-based approach is implemented for existing bridge inventories. The tables provided in
the Guideline provide the framework for including such analysis in the RBI methodology. These tables provide user guidance
for identifying appropriate inspection methods and/or NDE
technologies to address specific anticipated damage modes.
3.4.3.2Element-Level vs. Component-Level
Inspections
There exists under the current implementation of the NBIS
a variety of approaches to collecting, documenting, and storing data on bridge inventories within individual states. While
many states are licensed to use the PONTIS bridge management system, which is an element-level process for storing
inspection information and evaluating future programmatic
needs, the degree to which states fully implement the elementlevel inspection process varies. Other states use the componentbased system that is required under the NBIS; still others use a
span-by-span approach. However, to implement the RBI process, more detailed information than that typically required
for a component-based system is needed. A component-level
approach, which is intended to provide a single average or overall rating for the three major bridge components, does not provide sufficient data for assessing the likelihood of future damage
developing for most cases, and as such will not support an RBI
analysis. Information on the specific damage modes present on
the bridge, their location, and their extent are needed to assess
inspection needs. As a result, inspection needs under an RBI
process are more closely aligned with more detailed, elementlevel systems. The key characteristics that are needed to support
the RBI assessment are as follows:
Report the damage mode or modes affecting key elements

of the bridge,
Report the location and extent of the damage, and
Report on key damage precursors as developed through
the RAP assessment.
Precursors identified through the RAP process may include
evaluating specific elements of the bridge such as the joints
or drainage systems. Specific conditions that are precursors
necessary to assess the likelihood of damage in the future will
also be needed, such as the presence of rust-stained efflorescence or fatigue cracking. Many of these may be found in the
current AASHTO Bridge Element Inspection Manual (26), in

many cases as bridge management elements or defect flags.


The bridge management elements and defect flags may need
to be more fully developed under the RBI process as needs
develop for specific inventories.

3.4.4RAP
The RAP is an expert panel assembled at the owner level
to conduct analysis to support RBI by assessing the reliability characteristics of bridges within a particular operational
environment and the potential consequences of damage. The
performance characteristics of bridges and bridge elements
vary widely across the bridge inventory due to a number of
factors. Variations in the ambient environmental conditions
obviously have a significant effect, since some states have significant snowfall, and, as a result, apply de-icing chemicals to
bridges frequently, while other states are arid and warm, such
that de-icing chemicals may be infrequently or never applied.
Design and construction specifications vary between states.
Typical details such as drainage features, and use of protective
coating or other deterioration inhibitors, for example, sealers for concrete, vary between bridge owners as do traditional
construction practices, construction details, and materials of
construction. In terms of consequences, redundancy rules and
traditional policies vary somewhat between bridge owners,
with some bridge owners requiring four members to be considered redundant, while others require only three, for example. Owners may also have policies specifying girder spacing
or other configuration requirements. All of these factors
contribute to the operational environment of a bridge that
affects the likelihood and rate of deterioration of bridges and
bridge elements, and, to a lesser extent, the assessment of the
potential consequences of that damage. As a result, knowledge and expertise of the operational environment, historical
performance characteristics, bridge management and maintenance practices, and design requirements for bridges and
bridge elements are essential for conducting reliability-based
assessments.
The role of such expert knowledge of a specific operational environment is a typical component for reliability or
risk-based assessments of inspection needs. It is necessary
that individuals with historical knowledge of the operational
environment and typical deterioration patterns within that
environment participate in the process. This participation
is needed to effectively assess reliability characteristics of
bridge elements and to identify and prioritize key attributes
and factors that support the rational characterization of the
OFs and CFs. To utilize this expert knowledge, which is inherently local to a specific bridge inventory, a RAP is formed at
the owner level to conduct the reliability-based assessment.
TheRAP panel typically will consist of four to six experts
from the bridge-owning agency. This team should include

154

an inspection team leader or program manager that is familiar with the inspection procedures and practices as they are
implemented for the inventory of bridges being analyzed.
The team should include a structural engineer who is familiar with the common load paths and the overall structural
behavior of bridges, and a materials engineer who is familiar
with the behavior of materials in the particular environment
of the state and has past experience with materials quality
issues. Experts from outside the bridge-owning agency, such
as academics or consultants, may be used to fill technical
gaps, provide independent review, or simply supplement the
RAP knowledge base as needed. A facilitator may also be used
to assist in the RAP process.
3.4.4.1 RAP Expert Elicitation
Expert elicitation is a method of gathering insight into the
probability or likelihood of failure of a component, or of evaluating associated consequences when insufficient operational
data exists to make a quantitative, frequency-based estimate.
When failures are rare, or it is necessary to predict future failures, expert elicitation is used to provide quantitative or qualitative estimates (categories) for use in assessing inspection
needs or the likelihood of adverse future events. Processes for
expert elicitations are common in nuclear applications and
other safety-critical industries for performing risk assessments of operating events and assessing in-service inspection
needs (21, 22, 47, 48). Key elements of the elicitation process
include assembling appropriate subject matter experts and
framing the problem to be assessed for the experts in order
to elicit objective judgments. Consensus processes are used to
aggregate expert judgments and ensure contributions from
all of the experts involved (21). For RBI for bridges, expert
elicitation is used to:
Categorize the OF based on expert judgment:

Determine credible damage modes for bridge elements


and
Identify and prioritize key attributes that contribute to
the reliability and durability of bridge elements.
Assess likely consequence scenarios and categorize the CF.
The processes to elicit expert judgment from the RAP are
simple and relatively straight-forward. The primary purpose
of the processes is to provide a systematic framework that
allows for efficient, objective analysis, and allows for input
from all members the RAP. This allows for their expertise
to be utilized and for dissenting judgments or views to be
resolved such that issues are addressed as comprehensively as
possible. For example, to identify the credible damage modes
that are specific to the type of bridge and the element being
considered, the problem is framed for the panel by describ-

ing the element under consideration and its operational environment. The following question is then posed to the RAP:
The inspection report indicates that the element is in serious
condition. In your expert judgment, what is the most likely
cause (i.e., damage mode) that has produced/resulted in this
condition? This elicits from the panel a listing of damage
modes that are likely to occur for that element.
Each expert is asked to independently list the damage
modes he/she judges are most likely to have resulted in a failure of the element. The expert records each damage mode
and provides an estimate of the relative likelihood that each
damage mode would have resulted in the element being in
serious condition. The expert does this by assigning relative
probabilities to each damage mode, typically with a minimum precision of 10% (the sum of the ratings should be
100%). The expert may note supporting rationale for the estimate. The individual results from each member of the RAP
are then aggregated to evaluate consensus among the panel
on the most likely damage modes for the element. An iterative
process may be necessary to develop consensus on the credible damage modes for a given bridge element and identify
damage modes that are not credible. However, for many elements, the damage modes are well known and consensus may
be reached quickly.
Attributes are then identified through a follow-up process.
In most cases, the key attributes for a given damage mode can
be identified by posing the following question to the RAP:
Consider damage mode X for the subject bridge element. If

you were asked to assess the likelihood of serious damage


occurring in the next 72 months, what information would
you need to know to make that judgment?
This generates input from the RAP on what attributes of
the element are critical for decision making regarding future
expected behavior. The resulting input from the RAP can be
categorized appropriately and ranked according to the relative importance of the attribute for predicting future damage
for the identified damage mode and element. While there are
potentially many attributes that contribute to the durability
and reliability of a bridge element, it is necessary to identify
those attributes that have the greatest influence on the future
performance of an element. Rationale for each attribute is
documented, either by using rationale already provided in
the Guideline, or developing suitable rationale through a
variety of means including past performance, experience
with the given bridge element, input from the RAP members,
previous and contemporary research, analysis of historical
performance, etc.
Expert elicitation is also used for assessment of the CF by
providing different potential failure and consequence sce-

155

narios and asking the RAP to assign relative likelihood to the


outcome of the failure according to the CF scale. This is a useful tool for evaluating the appropriate CF for situations that
are not well-matched to the examples and criteria provided
in the Guideline, or to establish basic ground rules for the
assessment of common situations. The process involves a few
basic, but critical steps as follows:
1. Statement of the Problem: The RAP is presented with a
clear statement of the problem and supporting information to allow for expert judgment to be made. Care
should be taken to ensure the problem statement does not
contain information that could lead to a biased decision.
Theproblem statement typically includes data regarding
the bridge design, location, typical traffic patterns, and the
failure scenario under consideration.
2. Expert Elicitation: Independently, each member of the RAP
is asked, based on judgment, experience, available data,
and given the scenario presented, to determine the most
likely consequence resulting from the damage mode under
consideration. The expert is asked to express this as a percentage of the likelihood, with the smallest unit of estimate
typically being 10%. The experts may provide a statement
on what factors they considered in making the estimate.
3. Comparison of results: Once each member of the RAP
has rated the situation, the results of the elicitation are
aggregated. Generally, there will be consensus regarding
the most likely consequence. However, in some cases, the
most likely choice may not be clear and there will not be
consensus.
4. Identify CF: If there is consensus among the panel regarding the appropriate CF, then the rationale for making the
determination is recorded. This rationale should be consistent with criteria provided in the Guideline and if not,
the panel documents the deviation or changes and associated rationale.
For cases in which consensus is not reached in the initial
elicitation, the experts should discuss their rankings and their
assumptions and rationale for their specific judgments. The
members of the RAP should then be given the opportunity to
discuss the various judgments and to revise their scores based
on the discussion. In some cases, additional information may
be needed to support developing a consensus regarding the
appropriate CF. If consensus cannot be reached, a potential
approach would be to adopt the most conservative consequence scenario that was included among the revised scores.
Exceptions to the selected likelihood scenario should also be
documented.
The RAP may determine that additional analysis is required
to determine the appropriate consequence for a given damage scenario. In some cases, additional data collection may be

required in order to reach a consensus. Individual RAPs have


the flexibility to develop effective methodologies to address
cases in which consensus cannot be reached. However, the
method must result in the selection of the most appropriate
CF, based on the Guideline provided and sound engineering
judgment.
3.4.4.2 Example of Expert Elicitation
This section provides an example expert elicitation as an
illustration of the RAP process. As part of the research for
NCHRP Project 12-82, an expert panel was assembled of state
bridge engineers and inspection experts from seven different states and an engineer from the FHWA. The goal of the
two-day meeting was to have experts from several state DOTs
contribute to the development of reliability and RBI practices
for highway bridges by providing owner perspective on the
approach and tools being developed. The participants in the
meeting represented a good cross section of personnel from
state departments of transportation, ranging from personnel
responsible for overseeing bridge inspection activities at the
district level through the state-wide programs for inspection
and maintenance.
The meeting covered many of the topics necessary to operate a RAP at the state level, including identifying key damage
modes for certain bridge elements, identifying and weighting
bridge element attributes that contribute to the durability/
reliability of the element, and evaluating the consequences of
various damage modes. Among the activities at the meeting
was a trial of the suggested expert elicitations processed utilized in the Guideline for conducting the reliability analysis
needed as part of RBI practices. This section of the report
provides example results from this workshop to illustrate the
elicitation process and sample data provided by a cross section of practicing engineers. Although this panel included
individuals from a variety of operational environments, and
results of the elicitation process would likely have differences within a specific environment, the results are included
here to illustrate the process and provide typical results. The
example presented here includes the results for a steel bridge
superstructure. These same processes were used during RAP
meetings held as part of two case studies of the technology,
reported in Section 3.6.
3.4.4.2.1 Identifying Damage Modes. The process for
determining credible damage modes based on an expert elicitation was conducted during the workshop. The goal of the
exercise was to identify the most likely and credible damage
modes for the element and establish the consensus (or lack
of consensus) of the panel regarding the most common damage modes for that element. The panel was asked to perform
this assessment for a steel girder. The following question was

156
Table 6. Example of expert elicitation worksheet for steel girder
damage modes.
Damage Mode

Likelihood (in 10%increments)

Corrosion / Severe Section Loss


Fatigue Cracking
Impact Damage/ Fire
Overload
Stress Corrosion Cracking
10% 20% 30% 40% 50% 60% 70% 80% 90% 100%

posed to the panel, You are told a steel girder is condition


rating 3, serious condition, according to the current NBIS
rating scale. Based on your experience, what damage is likely
to be present? The expert was provided a form similar to
that shown in Table 6, except that the damage modes and
likelihood indicators were blank. Each member of the panel
completed the form, identifying the damage modes and relative likelihood with a precision of 10%. Table 6 illustrates
the results provided by one of the panel members. As shown
in the table, this member rated corrosion/section loss as the
most likely damage mode to be present, with fatigue cracking
and impact damage as less likely, and overload as a possible
damage mode. In this case, the panel member identified stress
corrosion cracking as a possible damage mode, but one that
was very unlikely such that no likelihood was assigned for
that damage mode.

Figure 4 shows the results from each of the panel members


for this elicitation exercise. It was the consensus of the panel
that the most common damage mode for a steel girder was
corrosion damage/severe section loss. This damage mode was
selected by everyone on the panel, typically with values of
greater than 50% likelihood.
The methodology is simple for many bridge elements for
which damage modes are well known, and it establishes the
consensus of the panel in regards to the most likely damage
modes. It also helps to identify damage modes that may be less
well known, but of concern for the particular state or bridge
inventory. For example, one member of the expert panel had
a different view of the most likely damage modes for a steel
girder, marking impact damage (40%) as the most likely
damage mode in his/her state. The particular state has large
areas of arid environment, and hence a different perspective

100%
90%

Likelihood of Occurrence

80%
70%
60%
Participant 1

50%

Participant 2

40%

Participant 3

Participant 4

30%

Participant 5

20%

Participant 6

10%

Participant 7

Participant 8

0%
Corrosion

Fatigue

Overload

Impact/fire

Damage Mode

Figure 4. Results of expert elicitation on steel girder damage modes.

157

on the most likely damage modes. This illustrates how RAPs


in different operational environments may identify and prioritize damage modes differently, depending on their operational environment and experiences managing their bridge
inventory. This is an advantage of the methodology, as damage modes that are most important to a given bridge inventory are identified through the process; these damage modes
are not necessarily the same across the diversified operational
environments of bridges across the country.
It should be noted that this type of expert elicitation is a
process for identifying and prioritizing likely damage modes
for a given bridge or family of bridges based on expert judgment. It is not necessarily repeated over and over again for
cases in which damage modes are well known. Rather, it is a
tool for establishing that there is agreement on the most likely
damage modes, capturing the expert judgment of the panel,
and ensuring that the analysis is comprehensive and considers all credible damage modes.
Through this process, damage modes for which the likelihood is very small or essentially zero can be sorted out from
more common damage modes through a rational process.
In most cases, as was shown here, likely damage modes are
expected to be well known by experienced bridge engineers
and consensus can be readily achieved.
3.4.4.2.2 Attributes. Once the primary damage modes
were identified, the panel considered the individual damage modes identified and the element attribute that contributed to the reliability of bridge element. For example,
for the damage mode of corrosion/section loss, the expert
elicitation consisted of posing the question to the panel,
For the steel girder, you are asked to estimate how long
it will be before significant corrosion/section loss would
occur for that bridge. What information would you need to
know to make that estimate? A group discussion was held
to identify and discuss the key attributes, and discuss their
relative importance to determining the future deterioration
pattern for the steel girder. The panel suggested that one of

the most important attributes was the maintenance cycle


for the bridge, or the maintenance activities that were typically performed as part of normal operations. This includes
such activities as bridge washing, cleaning away of debris
that may accumulate, and maintenance of joints. The consensus of the panel was that this was a highly important
attribute that should contribute to the rationale. The panel
also identified that the bridge deck type was an important
attribute that could potentially be a screening criteria for
those bridges that have, for example, open-grated or timber
decks. The panel identified that built-up members with the
potential for crevice-type corrosion, micro-environments
associated with traffic overspray, and condition history
(trend data) were other attributes that could be considered
in assessing the future performance of steel bridge elements
in terms of corrosion.
The attributes were ranked according to their importance
as high (H), medium (M), or low (L), and if the attribute was
potentially a screening criteria (S). Table 7 summarizes the
results of the discussion.
The attributes identified by a particular RAP in a specific
operational environment may differ from those indicated in
Table 7; however, these results are provided as an illustration
of the process of eliciting expert judgments from a RAP. Once
the attributes are identified and ranked appropriately, a simple scoring regime can be developed based on the results and
used to categorize the OF based on these attributes.
3.4.4.2.3 Consequence Scenarios. Expert elicitation to
determine CFs was also demonstrated. An overview of the
process for selecting the appropriate consequence category
for a given damage mode was presented to the panel. This
overview included several examples of different consequence
scenarios that might be experienced during the evaluation
process, and a review of the draft criteria for assessing the CF
within an RBI process.
An exercise was conducted to illustrate and test the use of
expert elicitation for evaluating the likelihood of different

Table 7. Summary of attributes identified by the expert panel


for steel superstructures.
Design Attributes

Loading Attributes

Attribute
Deck
Joints/Drainage
Built-Up Members

Rank

Attribute

Rank

Macro Env.

H,S

Micro Env.

Deck Type

M,S

Material Type

L,S

Age

Condition Attributes
Attribute
Existing
Condition
Joint Condition
Maintenance
Cycle
Condition
History
Trend

Rank

Debris Accum.

H,S
H,S
H
M

158

consequence scenarios. The purpose of this exercise was to


determine if, given a certain damage scenario, there could
be consensus on the most likely outcome of that damage,
based on the defined consequence scenarios and applied to
a specific bridge. This process can be used by an RAP to
develop and illustrate consensus and agreement with the
Guideline for assigning consequence categories, to address
situations that may not be sufficiently addressed or unclear,
or to address unique situations for which expert judgment
is required. The bridge presented to the panel was a multigirder steel bridge with an Average Daily Traffic (ADT) of
2000 and spanning a divided state highway. Photographs of
the bridge and descriptions of its structural configuration
were provided to each member of the panel in written form
for use in assessing different damage modes and associated
consequence scenarios.
Each panel member was provided with a handout that
included basic directions, a bridge description, photographs
of the bridge, and nine different damage scenarios to evaluate
independently. The panel members were asked to complete the
bubble chart for each damage scenario, as shown in Table8.
The results were collected and reduced to summary charts
showing the average assigned likelihoods.
There were nine damage scenarios presented to the panel,
ranging from fracture of primary member to delamination
and spalling of piers and abutments. The results of this exercise indicated that for certain scenarios, there was strong
consensus on the most likely consequence of the indicated
damage. For example, for the following scenario:
The overlay is debonding; approximately 20% of the deck
is spalled.
The assessment of the panel was distributed as shown in
the Figure 5. As shown in the Figure, there was consensus
from the panels independent assessments that this scenario
represented a low to moderate consequence. Discussion of
this scenario indicated that some panel members judged
that the consequences could be high, based on their interpretation of the failure scenario presented. Discussion of the
assessments quickly yielded assessment that the appropriate
CF was moderate.

Figure 5. Likely consequences of general deck spalling.

A second scenario of interest was a comparison between


fatigue cracking due to out-of-plane distortion vs. fatigue
cracking due to primary stress. For the former, the panel
rated the most likely consequence as moderate (~50% likelihood), for the latter, the most likely consequence as high
(>60% likelihood)for the multi-girder bridge utilized in
the exercise. It was also interesting that a scenario of one beam
fracturing was similar to a primary stress fatigue crack, >60%
likelihood that the consequence would be high according
to the consequence categories provided. Figure 6 shows the
average outcome for the fracture of one of the steel beams,
and as indicated in the figure, the elicitation indicated that
the most likely outcome/consequence for this scenario would

Table 8. Sample table for assessing


likelihood for damage scenarios.
Consequence
Category

Likelihood (%)

Low
Moderate
High
Severe

Figure 6. Likely consequences of beam fracture.

159

be high. These data, which illustrate the consensus formed


from the independent assessments of individuals from a number of different states, would be refined when applied within
a specific bridge inventory and operational environment.
A series of criteria and requirements are provided in the Guideline to assist in this process, and in many cases the CFs may be
governed simply by the criteria in the Guideline or owner policies regarding the treatment of redundancy or other factors. In
other cases, additional analysis or testing needed may be identified through the process. For cases not easily addressed or
well defined, this type of expert elicitation is especially useful
as a tool for developing rationale to support the categorization
of the CF or identified specific analysis needs.
These examples illustrate the process and feasibility of
expert elicitation for determining the key factors required
in the RAP assessment. The decisions regarding the likely
damage modes and potential consequences are very similar
to decision processes currently utilized by bridge engineers
to determine the urgency of repair needs, anticipate future
repair needs, and manage bridge inventories to ensure safety
and serviceability of bridges. These decision processes are
simply collected and aggregated systematically to provide
rationale for decision making regarding bridge inspection
requirements. Additional testing of the processes and evaluation of the consistency of the elicitation outcomes were conducted through case studies reported in Section 3.6.

3.5 Data to Support RBI Analysis


There are a number of resources available or that could
be developed to support the RAP assessment of the OF for
bridge elements by providing data to support decision making. While none of these sources for data provide perfect solutions, for example, for calculating quantitatively the OF, they
can provide data that supports decision making and rationale developed through the RBI process. This section of the
report describes a few of these resources, as well as important
consideration for utilizing these data for the reliability assessment of bridges. First, use and application of the qualitative
and quantitative data is described.

through specific probabilistic models, databases of failure


rates, or past performance data such as deterioration rate
models. These data are typically more in-depth and detailed
than qualitative data. This can provide valuable insight and
uniformity in approach, but developing such data can be
impractical for realistic situations that are too complex to be
modeled effectively. Data on past performance are frequently
incomplete or inaccurate, and in some cases can provide ineffective estimates of future performance (50). Additionally,
the effort required to collect and analyze the data may far
outweigh the value of the data in estimating future performance, particularly when the data are sparse, include a large
uncertainty, or design characteristics are evolving.
Qualitative data enables the completion of assessments in the
absence of detailed quantitative data. This qualitative data can
be augmented with quantitative data when and where available,
forming a continuum of data as shown in Figure7 (23).
The accuracy of results from a qualitative assessment depends
on background and expertise of the analyst (21). Quantitative
data, such as deterioration rate information measured from
NBI or bridge management software (BMS) data, can provide
supporting rationale for decision making, if handled appropriately. Estimates of precise numerical values (quantitative
analysis) can imply a higher level of accuracy when compared
to qualitative analysis, though this is not necessarily the case,
particularly when there is a high degree of uncertainty or
variation. It is the quality of the data that is most important
to support an analysis, and the fact that data are quantitative
does not necessarily mean they are more accurate. Difficulty
in effectively representing past experience, expert knowledge,
and bridge-specific conditions can result in quantitative
data that are biased and inaccurate, or whose applicability
to a specific situation is unknown due to a complex array of
assumptions utilized in developing the data. As a result, careful elicitation of expert knowledge from those most familiar with the operating environments, historical performance
characteristics within those environments, and the expected
future performance is used for RBI (21). Formal methods for
eliciting expert opinion for the purpose of risk assessment are
included in the Guideline and in the literature, as previously
discussed (47, 51, 52).

3.5.1 Quantitative vs. Qualitative Analysis


Industrial standards for reliability and risk assessment
recognize both quantitative and qualitative methods for
estimating the POF and consequences of failure. Qualitative
data typically are composed of information developed from
past experience, expertise, and engineering judgment. Inputs
are often expressed in data ranges instead of discrete values
and/or given in qualitative terms such as high, medium, and
low (although numerical values may be associated with these
levels) (21, 23, 29, 49). Quantitative data are data developed

High
Detail of
Analysis

Qualitative
Analysis

Semi-qualitative analysis

Low
Quantitative
Analysis

Figure 7. Continuum of data needed for qualitative


to quantitative analysis (23).

160

3.5.2 Data Needed for Assessment


To perform a reliability-based assessment, the primary
data required include data on bridge design characteristics
and details, materials, environment, and current condition.
Bridge inventory data describing the overall characteristics
of a bridge, such as can be developed from existing NBI data
tables, can provide some information. Data on materials and
design characteristics are generally available in the bridge files,
typical design and detailing practices, and local knowledge of
construction practices. Damage data describe the deterioration active or expected on a structure and estimate its effects
on the structure and rate of development. For damage data,
sources include general data available including the NBI
database, inspection reports and supporting data within a
DOT, element-level data for many states, and industrial data
such as the experience of bridge owners, previous research,
historical data, and historical experience. In some case, deterioration rate data or trends may be available and used as part
of the assessment of future performance. Additional data on
the anticipated performance of bridge elements is developed
through the RAP process based on expert judgment.
3.5.2.1Deterioration Rate Data
and Previous Failure Histories
Deterioration rate data such as that developed by Agrawal
(35) and others (40, 53, 54) can be used to support estimates
of future performance of bridge elements. However, there
are challenges to applying these data exclusively to determine
appropriate inspection strategies for bridges. First, data on
bridge deterioration is often not specific, expressed normally
in subjective condition ratings that may not capture specific
characteristics of the bridge or the deterioration mechanisms
that led to a certain condition rating. As a result, making
accurate predictions regarding future performance can be
challenging. Second, variation in the data is high, such that
estimating deterioration curves typically requires advanced
probabilistic analysis that develops mean estimates for the
population. These mean or average values provide information
on expected average performance of an overall population, but
not for a specific item within that population. Deterioration
rate data may need to be modified to adjust the data to local
operating and management conditions to be used effectively
to estimate the future performance of specific bridges or
bridge elements within a population.
However, deterioration curves and probabilistic failure
estimates are valuable to the RBI analysis process in several
ways. Deterioration curves can provide background and support rationale for engineering judgment regarding future
performance of bridge elements, based on past performance
when combined with an assessment of the key attributes for

the elements identified through the RAP process. If a bridge


owner had a population of bridge elements that were very
similar in design, and constructed at the same time and to
the same specifications and quality, and exposed to the same
environment, then accurate probabilistic estimates of future
performance could be developed. Generally, this would be
atypical of the bridge inventory. Consequently, the method
developed for RBI practices provides a means for incorporating such analysis, but does not rely on these data alone.
Considerations for utilization of deterioration rate data
include:
Similarity of operational environment: The RAP should

consider if the particular bridge under consideration shares


the same operational environment as the elements from
which data was obtained. Key elements of the operational
environment include the ADT, ADTT, macro-environment
of the bridges (severe environment vs. benign environment),
micro-environment (salt application, joint and drainage
conditions, exposure to overspray), and typical maintenance
and management (among others).
Similarity of Key Attributes: Key attributes that affect the
damage modes and mechanisms for the bridge element
should be similar for the bridge under consideration to
those from which deterioration rate data was obtained. This
may include materials of construction, design attributes,
and condition attributes. Quality of construction and years
in service may also be a factor.
Component ratings for superstructure, substructure, and
deck (and culverts) are provided for all bridges under the
NBIS scheme providing general information on the deterioration of the structural components over time, based on visual
observations. Element-level data are documented for states
using PONTIS or other element-level inspection schemes.
Obviously, these condition data are an important component
to evaluating the current condition of a bridge, at least in a
general way, and identifying bridges with low or high condition
ratings. These data can also be used to construct the deterioration curve data to support assessments, or to make estimates of
typical performance characteristics for bridges of a particular
design, as described below.
3.5.2.2 Inventory Data Analysis
Data from the NBI database can be analyzed to support the
rationale for bridge inspection intervals developed through
the RBI process. For example, historical NBI data can be analyzed to determine the average period of time a particular
bridge element remains in a certain condition rating. These
data can be utilized to support rational decision making and
the use of surrogate data, such as utilizing condition ratings

161

Figure 8. Graph showing Weibull distributions for time-incondition for prestressed bridges in Oregon.

of 7, Good Condition, as a surrogate for condition attributes


associated with a certain bridge type. For example, Figure 8
shows the time-in-condition for prestressed bridge superstructures in the state of Oregon. These data were developed
by examining 20 years of NBI data, and determining from
these data the time period (no. of years) individual bridges
remained in a certain condition rating, according to the
inspection results documented in the NBI. Weibull distributions were used to characterize the distribution of years
in rating for this population of bridges, and these Weibull
distributions are shown in the figure. Simply summing the
mean (average) number of years, historically, that a prestressed bridge has remained in each certain condition rating,
assuming a bridge component is currently rated a 7 and changes
to a 6 immediately, the average number of years to progress to
a condition rating of 3 is ~15 years. Given a maximum inspection interval of 72 months (6 years), at least two inspection
cycles would be completed within this 15year period. During
these inspections, if deterioration occurs more rapidly than
initially envisioned, the inspection interval is appropriately
reduced. These data support the rationale that significant
margin exist when considering a bridge currently in a condition
rating of 7. When considered within an RBI process, which
identifies attributes of bridges that are likely to cause more
rapid deterioration, such rationale is well-founded and based
on quantitative, historical data.
More complex analysis of such data may also be used,
including deterioration curves, probability calculations, etc.
BMS, such as the Pontis program, may provide data on transition probabilities or lifetime estimates based on Weibull
statistics, which can be utilized to provide quantitative data

to support the RAP analysis. These data can be used to


complement the RAP analysis. However, to effectively use
these data, information provided through the RAP process
is needed to ensure the relevance of the data as discussed in
Section 3.5.2.

3.5.3 Industry Data


The RBI practices rely on engineering judgment and experience with performance of engineered structures under
actual conditions to estimate future performance. So-called
industry data are developed from the existing body of
knowledge across the industry, frequently contained in the
body of research literature available, to inform and support
expert judgments. These data may include specific, quantitative data such as would be provided from models, or the combined or collective knowledge based on the existing body of
research and past experience across the industry. This section
provides two examples of industry data that can be used
to support analysis under the RBI process: a simple, commonly available modeling example and a collective knowledge example.
There exists a significant body of research concerning the
degradation of highway bridges by common deterioration
modes. There are two primary modes of deterioration that
cause bridge damagecorrosion of reinforcing steel in concrete, and corrosion of steel bridge components. Certainly
there are others, such as fatigue cracking, but corrosion and its
effects can be associated with much of the damage occurring
in bridges over time. Methods of determining the remaining life of elements and details based on fatigue mechanisms

162

are documented and well known. Because of the significant


importance of corrosion-based deterioration modes to the
degradation of bridges, there exists a significant foundation of knowledge regarding corrosion and its effects on
bridges, which can be leveraged to develop estimates of
future behavior based on the age, current condition, and
design attributes of a bridge. The rate of corrosion of steel
and steel embedded in concrete varies widely according to
localized conditions, with the local environment being a key
factor. Geographical regions where de-icing chemicals are
regularly applied generally have significantly higher corrosion rates than regions where de-icing chemical use is low or
even nonexistent. The local environment at the bridge, such
as leaking joints or poor deck drainage, also has a significant
effect. This section discusses generalized data regarding the
corrosion rates in steel, for both steel members and steel
embedded in concrete. This data is provided to illustrate
the type of industry data that can be used to support the
rationale used by an RAP during the assessment process,
and could be further developed if needed to address specific
situations, or utilized as current industrial knowledge for
general cases.
3.5.3.1 Corrosion in Concrete Structures
The rate at which corrosion damage may develop varies
widely for different geographical regions, depending on the
level of exposure of the concrete to corrosive agents such
as air-borne chlorides, marine environments, and the use of
de-icing chemicals. The main factors that contribute to steel
corrosion are the presence and amount of chloride ions, oxygen, and moisture. To illustrate how these factors affect structures located in different geographical regions, commercial
software was used to generate benchmark corrosion effects
models for different regions of the country.
One of the objectives of the modeling was to illustrate
the variation in the likelihood of corrosion damage occurring in different geographical locations across the United
States. Given that the inspection interval is uniform under
the existing system, and that corrosion presents one of the
most common and significant forms of damage to bridges,
this study was intended to examine how much variation there
might be in corrosion rates, and hence inspection needs, to
assess corrosion damage across the United States. The results
of the study are reported in terms of time to the initiation
of corrosion. The time to the propagation of damage varies
somewhat but can be considered to be on the order of 6 years
for uncoated reinforcement to 20 years for epoxy-coated
reinforcement, based on the rate that damage is expected to
propagate once initiated in the reinforcing steel (36). Design
parameters such as the amount of concrete cover, rebar spacing,
and concrete material properties obviously affect the rate at

which damage will propagate for a specific concrete component. These factors were assumed constant for the purposes
of evaluating how quickly the effects of corrosion might be
realized across different geographic regions.
Ficks second law of diffusion was used as the governing equation to account for differences between geographic
locations, such as temperature levels and ambient chloride
concentrations. Ficks second law of diffusion is generally
stated as:
d 2C
dC
= D 2
dt
dx
Where
C = the chloride content
D = the apparent diffusion coefficient
x = the depth from the exposed surface, and
t = time
The chloride diffusion coefficient, D, is modeled as a function of both time and temperature, which represents the rate
at which chloride ions travel through uncracked concrete.
Higher temperatures allow for an increase in chloride diffusion as the ions have more energy to move, as compared to
those in cooler temperatures.
For the modeling, the benchmark concrete mixture
assumed contained only Portland cement with no special
corrosion protection strategies. The value of 0.05 percent by
weight of concrete was used as the threshold chloride level
for corrosion initiation for the uncoated rebar. This was done
to represent a worst case scenario for corrosion initiation,
given that no corrosion mitigation strategies were employed.
Complete details on the analysis process are available in the
literature.
Six states across the United States that represented different geographical regions and thus different chloride build-up
rates on the surface of the concrete, resulting from chlorides
in the environment and de-icing chemical application, were
modeled. These states included Arizona, Arkansas, Florida,
New York, Washington, and Wisconsin. For each state, chloride diffusion rates were modeled for rural highway bridges,
urban highway bridges, and also for marine zones, where
appropriate. Cover depths of 1 inch and 3 inches were used
to illustrate the effect of concrete cover over the range of typical cover. Representative results of the analysis for an urban
highway bridges are presented here.
Figure 9 visually illustrates the difference in the modeled
time to corrosion initiation for different geographic regions.
As shown in Figure 9, there are vast differences in the model
time to corrosion initiation for different locations across
the country. For aggressive climates, such as New York and
Wisconsin, corrosion initiated in as little as ~7 years, while
in less aggressive environments, such as Arizona, corrosion

163

nale for a lower OF, or conversely a higher factor if they were


not used. Such modeling is relatively simple, widely available (the application used was available free-ware), and can
include other relevant attributes to provide quantitative data
to support the RAP assessment.
3.5.3.2 Corrosion in Steel

Figure 9. Time to corrosion initiation for different


states based on a diffusion model.

initiation is not anticipated for almost 70 years. While this


model does not consider localized effects, such as cracking
of the concrete that can greatly increase the rate of chlorides
intrusion into the concrete, it does illustrate that the time to
corrosion for a generic, uncracked case varies significantly
across geographic regions.
As shown in the figure, New York, Wisconsin, and
Florida have very similar behavior in terms of time to corrosion initiation. These environments would fall more toward
the severe or aggressive side of the exposure environment
scale. Washington falls within a more moderate exposure
environment. Arizona and Arkansas, with the slowest chloride diffusion rates, are more mild environments. What is
most notable in this data is that the time to corrosion for the
simple, benchmark situation varies over an order of magnitude across the different geographic regions modeled. This
data illustrates that given the important role of corrosion
in the time-dependent deterioration of bridges, uniform
inspection intervals are unlikely to be the most efficient solution to the inspection problem. Bridges in aggressive environments are likely to deteriorate more rapidly, and thus
require more frequent inspections than bridges located in
benign environments. This is only one among a multitude
of factors that contribute to the need for inspections; however, it is one of the most important and widespread. Data
such as those provided through this simple modeling can be
used, among other inputs, to provide supporting rationale
for categorizing the OF with the RBI system. Element attributes that contribute to increased corrosion resistance, such
as the use of epoxy-coated rebar or concrete mixes intended
to resist the effects of corrosion are also needed for the analysis. This is particularly true in aggressive environments in
which corrosion mitigation strategies might greatly increase
the time to corrosion if they were used, supporting ratio-

There is also a significant amount of available literature


related to the corrosion of steel bridges and the use and performance of protective coatings for steel bridge corrosion
control (37, 44, 45, 5557). During periods of active corrosion, it is generally accepted that corrosion rates of steels
under most natural exposure conditions follow a linear rate
to a point where the corrosion rate slows and flattens to a
steady state rate less than that of the initial few years of corrosion. During the initial stages of corrosion, the rust scale
builds up at the steel surface at a fairly consistent rate. Once
the scale covers the entire exposed surface in a uniform manner, the rate of corrosion is limited by the rate of oxygen diffusion through the intact rust layer. Although this pattern of
a deteriorating linear corrosion rate is dominant for boldly
exposed steel, the rate itself is highly dependent upon the specific exposure conditions. The corrosion rate tends to abate
over time for many environments, but for the most aggressive environments (marine) this reduction in corrosion rate
may not occur. Also, the corrosion rate at localized areas on
the same structure, or even the same steel member, can vary.
Therefore, it is prudent to view long-term corrosion rates
as maintaining a near linear corrosion rate over time and to
assume corrosion rates that are in the range documented for
steel exposed to high moisture, high chloride environments.
These corrosion rates tend to be in the range of 0.004 inches
to 0.006 inches per year, per side of exposed steel, and these
values can be used as a conservative planning rate to predict
the impact of corrosion on a deteriorating member. Because
of this relatively slow rate of corrosion section loss in steel,
the accumulation of damage in the near future is predictable,
particularly in a relatively short time frame such as the next
72months. The condition of the structural steel and protective coatings relative to corrosion can be easily assessed
during inspections. If the current condition is not well
understood, for example, the amount of section loss present
in the bridge is not known, an effective assessment may not
be possible. However, under an RBI scheme, the inspection
process to be used must ascertain the level of section loss
present, enabling the effective assessment of the likelihood
and severity of future damage occurring. This data provides
an example of the collective knowledge available and easily accessible that can be used to provide a basis for RAP
assessments.

164

3.6 Case Studies of the Methodology

3.6.1 Summary Overview of RAP Meeting

Two case studies were conducted to evaluate the effectiveness of the RBI method. The objectives of the case studies
were as follows:

The RAP meeting consisted of a series of designed expert


elicitations intended to develop comprehensive data models for RBI. Processes implemented during the case studies
were as described in Section 3.4. During the RAP, credible
damage modes pertaining to the family of bridges being
analyzed were identified through consensus of the RAP. Relevant attributes that contribute to likelihood of those damage
modesprogressing or occurring were also developed through
the designed elicitations. Following the identification of the
damage modes and relevant attributes, these attributes were
ranked according to their impact on the likelihood for that
damage mode (high, medium, or low) as a means of establishing an initial scoring approach. CFs for each damage mode
and bridge component are also developed through a designed
elicitation and consensus of the panel. Data from the RAP
meetings were subsequently analyzed by the research team,
organized into scoring models for each damage mode based on
the RAP results, and utilized in the back-casting procedure to
verify the effectiveness of the RAP results.

Demonstrate the implementation of the methodologies

with state DOT personnel and


Verify the effectiveness of RBI analysis in determining suit-

able inspection intervals for typical highway bridges.


To demonstrate the implementation of the methodologies
for RBI, two state DOTs were selected to be trained for and
execute an RAP analysis for a family of bridges in their states.
This included training in RBI technologies and executing
expert elicitation according the procedures described in the
Guideline. These RAP meetings resulted in data models for
determining the RBI requirements for a family of bridges.
These results were then tested to verify that the RBI practice
developed through the RAP process was effective in determining an acceptable inspection interval for the subject
bridges. This was achieved through a back-casting process
that utilized historical inspection records for specific bridges.
These inspection records were used to assess if the inspection
intervals identified through RBI would have been effective in
maintaining the safety and serviceability of the bridge, were
the RBI procedures applied in the past. This process compared the outcome of the RBI analysis with actual performance data for specific bridges, providing a validation of the
RBI approach.
The first case study was conducted for a sample of prestressed bridges in Oregon and the second one for steel bridges
in Texas. In each case a group of bridge experts were gathered
to conduct the RBI analysis during a 1.5 day RAP meeting
in the host state. The composition of the RAP panels consisted
primarily of state department of transportation engineers
involved in the inspection, maintenance, and management
of bridges within the state. The goals of RAP meetings were
to develop RBI practices for the subject family of bridges.
The objectives of the meeting were to identify and rank damage modes for each bridge component (deck, superstructure,
and substructure), discuss deterioration mechanisms that
lead to those damage modes, and identify related attributes.
These attributes were then ranked according to their impact
on the likelihood of severe damage occurring within a specified time interval. CFs associated with these damage modes
were also assessed.
This section of the report describes the outcome of the
case studies conducted in Oregon and Texas. This includes
an overview of the RAP meeting agenda, resources used in
the RAP meetings, and the results of back-casting completed
to verify the RBI approach.

3.6.2 RAP Meeting Attendees


The RAP meetings were attended by a variety of individuals from participating states, as shown in Table 9. The RAP
meeting in Oregon was attended by nine individuals, including DOT engineers, academics, and industrial representatives.
The industrial representative participating on the Oregon
RAP was from a fabricator that provided precast members
for bridge projects in the state. The RAP also included a university professor with active research in the area of bridge
evaluation and condition assessment. There were three individuals with Ph.D.s.
In contrast, the RAP in Texas was comprised of only five individuals, and all of the participants were employed by the Texas
DOT. The participants generally held Bachelor of Science (B.S.)
degrees, with one member holding a Master of Engineering
(M.E.) degree.
Most of the participants in the RAP meeting held at least
B.S. degrees in civil engineering. A little more than 70% of
the participants were registered Professional Engineers (P.E.).

3.6.3 Schedule and Agenda


The RAP meeting in each state consisted of a 1.5 day face
to face meeting in the host state. A webinar was presented
approximately 1 week prior to the RAP meeting, to familiarize participants with the overall process, field any questions participants may have, and identify any resources that
may be needed for the meeting. This teleconference consisted
of presenting overview slides introducing the concepts and

165
Table 9. Listing of RAP meeting attendees in Oregon and Texas.
Name

Emp.

Current Position

Highest
Degree

P.E.

Oregon
Participant A

Oregon DOT

Bridge Program Unit Manager

B.S.M.E.

Participant B

Oregon DOT

Structural Service Engineer

B.S.C.E.

Participant C

Oregon DOT

Senior Engineer

Ph.D.

Participant D

Oregon DOT

Bridge Operation and Standards


Managing Engineer

B.S.C.E.

Participant E

Oregon St U.

Professor

Ph.D. Str.
Eng.

Participant F

Knife River
Corp

Chief Engineer

Ph.D.

Participant G

Oregon DOT

Bridge Maintenance

Participant H

Oregon DOT

Bridge Planner & Financial


Analyst

M.S. of
Economics

Participant I

Oregon DOT

Senior Bridge Inspector

B.S.C.E., AE
Struct. Eng.

Participant A

TX DOT

Director of Field OperationsBridge Division

B.S.C.E.

Participant B

TX DOT

State Bridge Constr/Maint Engr

B.S.C.E.

Participant C

TX DOT

Senior Bridge Const. and Maint.


Engr

M.E.C.E.

Participant D

TX DOT

State Inspection Engineer

B.S.C.E.

Participant E

TX DOT

Bridge management Engineer

B.S.C.E.

Texas

approach of the research and the planned activities during the


RAP meeting. Most of the individuals that participated in the
RAP meeting also attended this webinar to be introduced to
the technology and prepare themselves for participation.
3.6.3.1 RAP Meeting Agenda
The meeting agenda was developed to establish an effective training pattern for the reliability assessment to be conducted. The previous expert panel meeting held during the
initial phase of the project acted as the model for the RAP
meeting agenda to be carried out in each state. However, in
developing the RAP agenda, it was decided that the training
goals would be best met by reorganizing the session into distinct training and execution phases. In other words, training associated with each of the aspects of the analysis, such
as CFs, OFs, etc. were provided for the entire process before
tasks to identify the parameters specifically for the family of
bridges to be examined in the case study. This is in contrast
to the expert panel meeting held during the initial phases

of the research, during which the elicitations for each factor


were conducted following training for that particular factor.
The primary motivation for this decision was to ensure that
the participants had a full and complete picture of how data
would fit together in the final analysis, before making any
decisions on what the parameters or attributes should be for
the particular family of bridges being analyzed.
The meeting began with an overview of the research
approach, describing the goals and objectives of the RAP of
the workshop and the overall research approach. This overview session was followed by a training session on how to
identify damage modes and attributes for bridge elements,
for the purpose of estimating the OF required for the analysis.
This session includes three exercises to illustrate the process
to be undertaken in the expert elicitation for identifying damage modes and key attributes, and ranking the importance
of those attributes in terms of the reliability of the element
under consideration. In these exercises, a typical two-span
steel bridge was presented as the example to pose questions
regarding the typical damage modes that would be anticipated

166

for this element. The members of the RAP recorded their


responses on the bubble sheets and subsequently discussed
the identified damage modes as a group. During these discussions, credible damage modes were identified for further
analysis.
This exercise was followed by an elicitation of attributes
related to the reliability/durability associated with the primary
damage modes identified by the group and prioritization of
those attributes from high to low. This exercise illustrated the
process of the developing attributes and a semi-quantitative
scoring scheme for a particular family of bridges as a means
of identifying the OF for the RBI analysis. The process illustrated in this example is later repeated by the RAP for the
superstructure, substructure, and deck components for the
subject family of bridges (i.e., prestressed superstructures in
Oregon and steel superstructures in Texas).
Training was also provided on the CF categories that are
part of the analysis. A group exercise expert elicitation for
consequences was administered to illustrate the process of
identifying a consequence ranking for a particular damage
mode scenario. During this exercise, panel members considered the likely consequences of an identified damage mode
progressing to the defined failure state (e.g., serious condition) in terms of safety and serviceability of the bridge.
Following these exercises, the expert elicitation for the family of bridges under consideration was conducted. Separate
sections of the meeting address the superstructure, substructure, and deck components of the bridge. The same process
implemented in the illustrative examples was conducted for
each component to identify the likely damage modes, attributes contributing to the reliability considering those damage modes, and prioritization of the attributes. These data
were used to identify criteria and develop the initial scoring
scheme to be implemented for assessing the OF for the various damage modes identified through the process.
Consequence scenarios for each damage mode were also
developed through group discussions. During this task, each
damage mode identified in the earlier exercises was considered, and an expert elicitation was conducted to identify the
appropriate CF for each damage mode, and key factors that
affect the factor selected. For example, if the damage mode
is spalling damage on a deck, the CF may be high or even
severe if ADT and traffic speeds are high, but moderate if the
ADT and traffic speeds are low. Group discussion was used
to develop consensus on these factors. Policies and common
practices in the particular state also contributed to these
discussions.
The balance of the agenda was used to refine and complete the criteria and rankings for attributes, OFs, and CFs
for the subject family of bridges. Screening criteria, surrogate
data, and available data on attributes from existing inspection
practices were identified. For example, if the subject state col-

lects element-level data, how do various element ratings and


damage flags correspond to the attributes and damage modes
identified through the RAP process.
At the completion of the meeting, it was anticipated that
the damage modes, ranking for attributes, and basic scoring
approach would be completed, as well as the CFs for various
scenarios. However, discussions of the CFs revealed that certain descriptions of the various CF levels were problematic,
and these descriptions were subsequently modified to address
these concerns. As a result, the RAP meetings provided preliminary data on the CFs to be used for the analysis, and these
were later refined during the analysis process.
The data from the RAP meeting were compiled and analyzed by the research team following the meeting. These data
were utilized to developed scoring models, or data models,
reflecting the input from the RAP. These data models were
then used in the back-casting process to evaluate the historical performance of a sample population of bridges in each
state to verify the effectiveness of the data models developed
through the RAP process.
3.6.3.2 RAP Participants Notebook
A participants notebook was prepared for distribution
to members of the RAP. This notebook provided a reference for use during the meetings. This notebook included
standard information regarding the meeting, such as the
agenda and copies of the slides to be presented during the
training portions of the meeting, including space for participants personal notes. In addition, copies of the forms to be
completed during the meeting are included for future reference following the meeting. The notebook also included
color copies of the risk matrices to be used in determining
the inspection interval based on the RBI analysis conducted
by the RAP.
The notebook also included key appendices from the Guideline. These appendices include the guidance for identifying
damage modes and attributes (i.e., OFs), CFs, determining
the inspection interval, and the complete index and commentary of attributes identified in the Guideline. These portions of
the handbook were included to act as references for the RAP
participants to use during the RAP meeting for conducting
the RBI analysis.
3.6.3.3 Software Development
A software application was developed to support the RBI
analysis of bridges based on the results of the RAP meetings.
This application was developed within a spreadsheet program, and provides a simple and rapid means of implementing the damage modes, attributes, and scoring methodology
for estimating the OF.

167

Figure 10. Example screen from software


application showing selection of attributes.

In this application, the user selects the attributes identified


by the RAP for a particular damage mode, as shown in Figure 10. A check box is used to select screening, design, loading, and condition attributes as described in the Guideline.
Reserved attributes are included so a user can easily add additional attributes that may not be included in the Guideline.
Once the attributes are selected from the appropriate listing, the application organizes the selected attributes into a
scoring page as shown in Figure 11. On this screen, pulldown menus are used to score the individual attributes for a

particular bridge according to the scoring scheme developed.


These pull-down menus allow a user to quickly select the
appropriate ranking for a particular attribute based on the
criteria developed through the RAP.
The individual scoring for any attribute can be easily
modified on an editing page to meet the requirements of a
particular user. A hot-link is provided to the attributes commentary included in the Guideline, such that a user can easily
refer to the rationale for a particular attribute and the envisioned scoring mechanism. After each attribute is scored, the
OF score and guidance is automatically calculated for that
damage mode.
This software application was developed for use in the case
studies to implement the analysis of the RAP from each state,
and for testing that analysis against the historical performance of bridges during the back-casting. Looking forward,
this software application provides a model for future, more
sophisticated computer applications to allow for efficient and
simple application of the RBI technology. For example, such
a software module could be an add-on to the PONTIS program or other BMS, where many aspects of the scoring could
be automatically obtained based on element ratings already
collected as part of a routine inspection.

3.6.4 Back-Casting Procedure


The case studies conducted in Texas and Oregon developed a set of criteria and attributes for determining the OF
and the CF, resulting in inspection intervals based on the risk
matrix. These criteria and attributes produced a risk-based

Figure 11. Example screen from software application showing pull-down menus for scoring attributes.

168

data model to be used to determine the appropriate maximum inspection interval for a specific bridge or family of
bridges. To verify if the use of these models provided a suitable inspection interval that did not compromise the safety
and serviceability of bridges, a back-casting procedure was
used. In the back-casting procedure, the data models developed by the RAP were applied to individual bridges based
on historical inspection records. For example, the data model
may be applied to a bridge based on the year 2000 inspection
records for the bridge, resulting in an RBI interval that would
have been determined in the year 2000, were RBI practices
applied at that time. These results were then compared with
the actual performance of the bridge, based on the inspection
records for the years 2002, 2004, 2006, etc. to determine if the
RBI inspection interval would have adequately addressed the
inspection needs for the bridge. The criteria for determining
the effectiveness of the data model included:
1. Did the condition rating for any component change significantly during the RBI interval in a manner that was not
captured or anticipated effectively, but would have been
captured (or detected sooner) by a standard, 24-month
interval?
2. Were there any significant maintenance or repair actions
completed that would have been delayed as a result of implementing an RBI interval (relative to a standard, 24-month
interval)?
3. Were there any significant factors or criteria not identified through the RAP analysis that were needed in the data
models to provide suitable results?

The procedure for back-casting consisted of obtaining the


element-level inspection reports ranging back to approximately 1998, depending on the availability of data for the specific bridge. The data model was applied at each inspection
year to assess the appropriate inspection interval based on the
inspection data. As a result, the RBI interval may be consistent over the time period examined, decrease over that time
period, or even increase during the time period as a result of
a repair or improved condition rating or condition state.
The overall concept of back-casting is shown schematically
in Figure 12. This figure shows NBI ratings for an example
bridge component over time. The RBI data model is applied to
the bridge component based on inspection results from 1998.
Assuming this results in an inspection interval of 72months,
the inspection results from each biannual routine inspection
(24 months) is examined to see if there were any significant
changes to the condition, or other events or circumstances
detected by the routine inspection that may have been missed
or delayed due to the RBI interval of 72months. The RBI interval is calculated for each year there is an inspection result, indicated by the numerical results shown on the diagram. A change
in the RBI inspection interval to 48months is also shown in
the figure. Assessment of the results includes determining if
the change of inspection interval identified through the RBI
criteria was effective in capturing the appropriate inspection
interval, considering changes in the condition of the component reflected in the inspection results. It should be noted that
the RBI inspection interval does not necessarily reflect NBI
condition rating changes; however, since both depend on the
condition of the component, they may be similar.

Figure 12. Graph of condition ratings for a bridge component


over time, showing schematic example of the back-casting
procedure.

169

3.6.4.1 Inspection Data for Back-Casting


Inspection data from each state were reviewed in detail to
implement the data models developed through the RAP process, i.e., evaluation of the attributes identified by the RAP.
This included design and loading attributes, which typically
do not change over the life of the bridge and condition attributes that change as the bridge ages or undergoes repair or
rehabilitation.
Inspection data from Oregon consisted of PONTIS data
file outputs, including photographs, notes, and standard
Structural Inventory and Appraisal (SI&A) sheets. Available inspection data from 1997 to present were assessed. In
Texas, inspection data consisted of inspection reports that
included standard SI&A sheets, NBI component rating sheets,
and element-level data collected at the time of the inspection.
Available data from 1999 through present were assessed for the
Texas case study.
3.6.4.2 Review of Work
As part of the back-casting analysis, a database of work
projects maintained by the Oregon DOT was queried to
determine if any significant repairs had been completed on
the subject bridges during the interval of the back-casting
process, and results were provided to the research team.
This was done to ensure no significant events occurred on
the bridge that resulted in major work or repair between
inspection intervals, which may have been missed due to
an extended inspection interval or may not be reflected in the
inspection reports.
In Texas, inspection records were more diversified. Work
during the intervals between inspections was determined from
the element-level data collected as part of the bridge inspection
process. This was achieved by reviewing each inspection report
for notes that would indicate that an improvement or repair
was made to the bridge, or that an improvement or repair
was urgently needed. Unexplained changes in the condition
rating for a component were also investigated to determine if
an urgent repair or rehabilitation activity was the source of the
improvement.
3.6.4.3Sampling
To complete the back-casting verification study of the
result of the RAP assessment, a sample population of bridges
was assessed over a time period dating back 15 to 17 years. To
determine the number of bridges to be assessed to develop a
statistically significant result, a statistical analysis of population sampling was completed. Generally, such statistical
models require some a priori knowledge of the anticipated
variance in the population to be sampled to estimate the

number of samples required to represent the overall population, considering the parameter to be measured. It was
anticipated that the RBI criteria developed by the RAPs would
include the current condition rating for a bridge as one of the
criteria (attributes). Therefore, it was desired to select a sampling of bridges that has the same variation as the population
overall, namely, that the natural variation of the inspection
results of the overall population is represented in the sampling selected, based on the condition ratings provided in the
inspection files. Experimental data from the FHWA visual
inspection study (58) was used as a basis for the estimate,
assuming that the variance of condition rating for all components in the FHWA study. Based on population sampling
statistics, assuming that the desired accuracy was 0.5 condition ratings with 95.5% confidence resulted in a desired
sample size of 17 bridges. For a confidence interval of 95%,
the sample size for back-casting would be 10 bridges. Based
on these results, the sampling of bridges included a minimum
of at least 10 bridges; in the study, 17 bridges were selected
from Texas and 22 bridges were selected in Oregon.

3.6.5 Statistical Analysis of NBI Data


Statistical analysis of NBI data for the participating states
was conducted to identify the characteristics of the each
states inventory and to support the RBI analysis. Analysis of
NBI data was completed with the following objectives:
1. To determine the typical characteristics of the bridge
inventories in the participating states of Texas and Oregon.
2. To develop quantitative data based on NBI condition rating
history to be used to support the RAP analysis and rationale for RAP-developed criteria.
The objective of providing quantitative statistical data
to support anticipated criteria that may be developed by the
RAP during the course of the case studies can be illustrated as
follows. Consider that the RAP identifies an attribute/criteria
(among others) that a bridge has a superstructure rating of7,
based on the rationale that such a condition rating would
indicate little deterioration or damage presently, and a low
likelihood (i.e., OF) that severe damage would occur over the
ensuing 72-month period. Analysis of the time-in-condition
data from the NBI records provides quantitative data to support this rationale, as discussed in Section 3.5.
To conduct these analyses, data from the NBI dating back
to 1992 were obtained from the Federal Highway Administration (http://www.fhwa.dot.gov/bridge/nbi.cfm); these
data are publicly available via the web site indicated. These
data were used to develop data on the past performance of
bridges in each of the participating states and to characterize
the overall inventory in each state.

170
Table 10. Bridge population statistics for Texas and Oregon.

Description
Concrete

Bridge Inventory in Oregon


Length
% of
% of
No.
(m)
No.
Length
2,050
87,000
28
24

Average Age (year)


55.2

Steel

1,089

109,160

15

30

48.5

Prestressed concrete

3,612

154,877

49

43

28.9

Other

602

12,408

49.3

Total

7,353

363,444

100

100

40.8

Concrete

29,098

704,514

56

23.40

48.0

Steel

7,423

776,717

14

25.90

38.1

Prestressed concrete

13,781

1,392,706

27

46.30

23.6

Bridge Inventory in Texas

Other

1,576

131,465

4.40

33.0

Total

51,878

3,005,403

100

100

39.6

3.6.6Bridge Inventories in Texas


and Oregon
The families of bridges selected for the two participating states were based on the bridge inventories in each
state. It was as desirable to have a sufficient inventory as to
have a large inventory from which to draw sample bridges,
and a representative population of bridges in terms of age.
Table10 shows bridge population statistics for the participating states based on data available in the NBI. The bridge
families selected are highlighted for both Oregon and Texas.
Prestressed bridges were selected for analysis in Oregon because
this bridge superstructure type made up almost 50% of the
bridge inventory in that state, making it a significant population of bridges. This population of bridges has an average
age of almost 29 years, consistent with the era of prestressed
bridge construction, and there are more than 3,600bridges of
this material type. In Texas, the overall number of bridges is
large, such that any family of bridges of similar superstructure
materials would provide a suitable population of bridges for
analysis. In this case, steel bridges were selected for analysis for
three reasons; first, they provided suitable number of bridges
for analysis, second, it was desirable to do one analysis for concrete and the other for steel bridges, and, finally, the average
age of the population was much older than the prestressed
bridge population in Oregon, providing diversity in the ages
of populations in these states.
Figure 13 illustrates the age distribution for bridges in each
state, as well as the age distribution for the bridge sample
selected for analysis. Vertical lines on the figure indicate the
mean ages for each population. As these distributions illustrate, the mean or average age of bridge selected for analysis
were older than the overall populations. This was considered
desirable, because relatively new bridges are generally less
challenging for RBI analysis, because they are usually in good
condition and have good durability attributes. Therefore,

selecting a population that was slightly older than the overall


population presented a greater challenge for testing the RBI
processes.
Bridges included in the sample were generally randomly
selected, with the exception that the desired sample of bridges
for analysis had a geographic distribution across the subject
state, and emphasis was placed on including bridges with sufficient historical data to make the back-casting meaningful.
In Oregon, several bridges had limited historical data because
the bridge was constructed after the year 2000; however, the
sample of bridges was larger such that there were at least
17bridges with the desired historical data available.
3.6.6.1 Bridge Sample Locations
Bridges selected for back-casting were distributed geographically within the states. Figure 14 shows the distribution
of bridges in each state. As shown in the figure, bridges were
selected from different regions of each state, although the
geographic distributions of the sample bridges are affected
by the population characteristics of each state. For example,
in Oregon, population density is significantly higher in the
western part of the state, and as such, the majority of bridges
are in the western part of the state; the sample of bridge
reflects this effect.

3.6.7 Time-in-Condition Rating


The NBI data for Texas and Oregon were analyzed to determine the typical lengths of time that a bridge component was
in a particular condition rating. These data were derived from
the NBI database, with some data trimming to accommodate
the fact that the data sets are incomplete. That is, there are no
data prior to 1992 or after 2011, so some trimming of these
data are needed to improve the certainty of the derived time
intervals. Data were trimmed from the data set if there were

171

Figure 13. Age distributions of sample bridges and overall


populations for (A) prestressed bridges in Oregon and
(B) steel bridges in Texas.

5 years or less of consecutive data in a condition rating at


the beginning or the end of the available time interval. The
trimming value of 5 years was selected based on study of different possible trimming values, ranging from 3 to 7 years,
performed by the research team. This study indicated that
the specific trimming value only had modest effects on the
outcome of the analysis, and as such, 5 was selected as an
acceptable value that ensured sufficient data were available
for a meaningful statistical analysis. This method of trimming the data provides a suitably conservative result, because
the analysis indicates that time-in-condition ratings are typi-

cally much larger than 5 years for components in reasonably


good condition (rated 6, 7, or 8). Data presented within this
report include the superstructure and deck condition ratings;
data for substructures were also analyzed. However, the deck
and superstructure condition ratings typically change more
frequently than substructure ratings, and as such, the deck
and superstructure are the focus of the data reported herein.
Figure 15 shows the time-in-condition results for prestressed bridges and decks of prestressed bridges in the state of
Oregon. As shown in the figure, bridge superstructures rated in
good condition tend to have longer intervals in that rating; as

172

Figure 14. State maps showing


geographic distribution of
sample bridges.

the rating decreases, the time in a particular rating is reduced.


For example, for the prestressed bridge superstructures illustrated in the figure, the average time period a superstructure
was rated 8 was almost 14 years (s = 4.9years), but the
time period a superstructure is rated a 5 is less than 5 years
(s = 2.7 years). The average time period for a prestressed superstructure is rated a 6 is 6.5 years (s = 3.8 years). For bridges in
condition ratings of 4 or 3, these data are not particularly useful
for two reasons; first, there are very few bridges in this category, and second, the bridges get repaired, and as such, the
time interval in the condition is really more representative of
a measure of how quickly these bridges may be improved or
repaired rather than how long they remain at this condition
rating.
Figure 15B shows the time-in-condition rating for decks
of bridges with prestressed superstructures. Similar observations can be made, as shown. For example, a deck remains
in condition rating of 7 for 10.2 years (s = 5.03years); the
time period a 6 remains a 6 is 6.4 years (s =4.8 years), on
average.
Figure 16 shows the results of the trimming analysis for
steel bridges in Texas. In this case, steel superstructures and
bridge decks on steel superstructures were analyzed. For
steel superstructures in Texas, the average time-in-condition
rating of 7 was 10 years (s = 5.4 years), for decks of steel
bridges, the average time-in-condition rating was found to
be 11 years (s = 5.6 years).
These data are useful as they reinforce and support the
supposition that a bridge in good condition tends to stay in
good condition for a long time interval (i.e., longer than the

Figure 15. Time-in-condition rating for (A) prestressed


bridge superstructures and (B) decks based on NBI
data for Oregon.

maximum inspection interval recommended using the proposed methodologies). For example, if one used the surrogate
data of condition rating of 7 for superstructure, substructure,
and deck to identify bridges with an appropriate inspection
interval of 72 months, these data provide quantitative evidence
to support that rationale, as discussed in Section 3.5. These
data were used in the case studies to support surrogate data
analysis based on the data models developed by each RAP.
In this analysis, the condition rating of 7 was used as surrogate data for the condition attributes to assume the OF
would be low for condition-related damage modes. For these
cases, the inspection interval of 72 months may be applied,
based on these data.
3.6.7.1 Inspection Intervals
Inspection intervals were determined based on the reliability matrix introduced in the Guideline. Figure 17 shows
the proposed reliability matrix that is used for typical
highway bridges. This matrix illustrates the appropriate

173

CF for the given bridge component and damage mode. Each


damage mode for each bridge component was analyzed using
the RBI procedure, resulting in a data pair (OF, CF) for each
damage mode for each component. These data were located
on the risk matrix to determine the inspection interval for
each bridge, as illustrated in the results section for each state.

3.6.8 Overview of Case Study Results


The objective of this section of the report is to provide an
overview of the results of the RAP meetings in each state. This
section includes a summary of the damage mode and attributes identified in each state, and the consequence analysis
that was conducted during the RAP meetings.
3.6.8.1 Summary of Damage Modes and Attributes
This section summarizes the damage modes and attributes
identified through the RAP process. These data provide the
data model for assessing the OF as part of the RBI process,
and as such, are documented here to illustrate how the data
model was developed and what was considered. Due to the
detailed nature of many of the attributes and description,
most of these data have been placed in Appendix A for the
Oregon case study and Appendix B for the Texas case study.
These appendices document the attributes and attribute
scoring for each damage mode that was used during the backcasting analysis.
Figure 16. Time-in-condition rating for steel bridge
(A) superstructures and (B) decks based on NBI data
for Texas.

inspection intervals based on the estimates of the OF and


the CF from the RAP analysis. In the figure, the inspection
intervals are I =12 months, II = 24 months, III = 48 months,
IV = 72months, and V = 96 months. For example, when an
OF is Low and CF is High, the proposed inspection interval is 48 months. This matrix was applied to the results of the
OF analysis, based on attribute scoring, and the appropriate

Figure 17. Reliability


matrix for RBI.

3.6.8.2 Damage Modes and Attributes


The expert elicitation process described in the Guideline
and implemented during the case studies generally worked
effectively to ascertain credible damage modes and identify
key attributes affecting those damage modes. The process
consists of having participants complete forms identifying
credible damage modes, and then using a consensus process
to list the damage modes, pare down those that are repetitive or irrelevant, etc. During the consensus process, damage modes identified by participants were recorded on a
white board, along with the data from the likelihood estimates made by the participants. An example of this process
is shown in Figure 18. This figure illustrates the beginning
of the expert elicitation process, when the data from each
member of the RAP is collected for discussion. The orange
numbers shown in the figure indicate the number of panel
members recording a particular likelihood (10%, 20%, 30%,
etc.) for a given damage mode. As shown in the figure, this
initial process included a number of damage modes for decks,
including rebar corrosion, delamination, and spalling, which
were pared down through discussion to a corrosion-related
damage mode of spalling.

174

Figure 18. Example of RAP data for damage


modes in decks.

Rutting was also identified as a credible damage mode for


decks by the RAP in Oregon. This damage mode illustrates
one benefit of a RAP consisting of bridge owners. Rutting of
decks is related to the use or over-use of studded tires, and
occurs along particular corridors in Oregon. It is unlikely
many other states would identify this damage mode, but
regardless, in Oregon such damage occurs and affects the serviceability of some bridges. It was the consensus of the panel
that this damage mode was credible and required consideration in an RBI process.
Texas identified punch-through as a credible damage mode.
In this case, punch-through is not a corrosion-related damage mode, but rather related to the construction of thin decks,

sometimes with poor quality concrete, that results in punchthrough as a result of repetitive loading and age. Because much
of the state is relatively arid, and use of de-icing chemicals is
minimal, decks may have longer lives than they might in an
area where corrosion is a significant issue. If the deck is thin
and concrete quality is poor, punch-through can occur. Like
rutting, this damage mode is due to local (state) policies and
construction practices, namely that very thin decks were used
during certain historical time intervals, and concrete quality
was not well controlled at the time. In a state where corrosion
damage was more prevalent, such a deck would deteriorate
severally due to corrosion before such punch-through could
occur. Like rutting, this damage mode is not likely common
in other states. These relatively unique damage modes illustrate the utility of the RAP approach.
A summary of identified damage modes is shown in Table11
for Oregon and Texas. It can be seen that damage modes of
concrete deck and substructure are similar for Oregon and
Texas. For superstructures, only the impact damage mode was
common between prestressed and steel bridges analyzed in the
two states, as would be expected, since the superstructures are
of different material types.
During the Oregon RAP, the panel expanded its assessment from open prestressed shapes, such as typical
AASHTO shapes and Bulb-Tees, to include adjacent box
girders bridges and prestressed slabs. The consensus of
the panel was that the damage modes and attributes were

Table 11. Summary of damage modes in Oregon and Texas.


Bridge Element
Deck

Oregon (Prestressed or Post-Tensioned)

Spalling
Rutting
Cracking (Non-corrosion
Induced)
Cracking (Shear)
Strand Corrosion
Fire Damage
Impact
Rebar Corrosion within the Span
Bearing Seat Problems

Superstructure

Adjacent Box Girders


Rebar Corrosion/Section Loss
Strand Corrosion (Fracture)
Flexural Cracking
Shear Key Failure
Impact / Fire

Substructure

Settlement
Corrosion Damages
Fire
Overload Damages
ASR

Texas (Steel)

Spalling
Punch-Through
Cracking
Delamination

Fatigue Cracking
Section Loss
Fire Damage
Impact
Deflection Overload
Bearing Failure

Settlement
Corrosion Damage
Overload Damage
ASR

175

element-level manual. In other words, the RAP identified


a given attribute and appropriate scoring limits during the
meeting, and these were later found to match existing element descriptions in the Oregon element manual. A similar
process was followed for Texas, which collects more limited
element data during inspections.

essentially identical for these families of bridges, with the


exception that adjacent box girder bridges had a shear key
damage mode that would need to be assessed as a screening tool. These data are reflected in the summary of damage
modes shown in Table 11.
For each damage mode identified by the RAP, attributes
that contributed to the likelihood of that damage mode
occurring and progressing were identified through the RAP
survey process and consensus of the panel. An example result
of the consensus process is shown in Table 12 for deck spalling, summarized from the Oregon RAP. In this table, the
attributes identified by the panel are shown in the left column, followed by the rank that each attribute was assigned
by the panel. This rank shows the unanimous vote on the
rank for each attribute; this represents consensus developed
among the panel, not necessarily initial results of the elicitation process. In some cases, individual members may have
ranked these attributes differently, but consensus was developed through discussion. Once consensus was developed,
limits or parameters for scoring each attribute was developed
through open discussion among the RAP member and results
are shown in the table. For Oregon, which utilizes elementlevel inspection processes, many of the attribute parameters
could be described using existing models from their elementlevel inspection manual.
For example, for deck cracking, the element manual
already has quantitative description of condition states 1, 2,
3, and 4, and therefore additional description was not necessary. For other attributes, for example ADTT, limits for high,
medium, and low were developed through discussion. A
comprehensive listing of the damage modes, attributes, and
limits/parameters used in the back-casting analysis are
included in Appendix A. The potential source of the data,
based on state-specific inspection processes, is also tabulated in AppendixA. It should be noted that in some cases,
the RAP identified attributes that were later correlated with
existing element data following a more detailed review of the

3.6.9CFs
Designed expert elicitations were also used to develop
CFs for each of the damage modes during the RAP meetings.
For most damage modes, singular failure scenarios were
assessed for each bridge component. The failure scenarios
considered consisted of the component condition rating
being serious (CR = 3), not necessarily structural failure.
For decks, for example, the scenario considered in that the
deck deterioration would typically be considered serious
(CR = 3) during a normal inspection. For superstructure
components (i.e., prestressed girders or steel girders), loss of
load-carrying capacity for one member was considered. For
Oregon, the CF for deck damage and substructure damage
was considered to be generally Moderate. For superstructure components, the initial CF developed in the RAP was
High for most damage modes (except bearing area damage);
this factor was subsequently discretized during the analysis
process.
For Texas, issues were identified during the RAP meeting with the CF descriptions, as previously described, and
these CF descriptions were subsequently adjusted during
the back-casting to address these issues. These revisions
adjusted the descriptions of different CF levels, but not the
levels themselves.
The CFs were subsequently assessed during the backcasting according to a series of scenarios to test and evaluate
the influence of different parameters on the analysis. These
focused largely on the CF assigned to the bridge superstructure. The scenarios included considering the CF as uniformly

Table 12. Example attributes rankings for deck spalling from the
Oregon RAP.
Attributes
Cracking
Delamination
ADTT
Location /
Environment
Age
Dynamic Loading
Rebar Corrosion
De-icing

Rank
H M L

8
8

>25%
>5000

8
8

Coastal and Mountain


8

>50

8
Rust/Black/Low
Cover

8
8

High

Limits
M
Existing model
11%-24%
501-4999
Valley (general
environment)
10-49
Existing Model

L
<10%
<500
Desert
<10
No stains,
Epoxy/high
cover
Low

176

High and considering the CF as uniformly Moderate, or determining the CF based on structural redundancy and feature
under the bridge. For the latter, the CF was based on the following criteria:
The CF was Moderate for the superstructure if:
Superstructure consisted of more than four members AND
Beam spacing of 10 ft or less AND
Bridge not over a roadway.

The CF was considered High if:

3.7 Back-Casting Results for Oregon


This section summarizes the results of the back-casting
analysis conducted as part of the study. The State of Oregon
provided 22 bridges from around the state for the analysis, as
shown in Figure 14A. As shown in this figure, bridges were
obtained from across the state to represent different environmental conditions surrounding the sample bridges. The
damage modes, attributes, and data scoring models used in
the back-casting process are documented in Appendix A.

3.7.1Environments

Superstructure consisted of four members or fewer OR


Beam spacing was greater than 10 ft OR
Bridge was over a roadway.

These criteria were based in part of the result of previous


NCHRP research on redundancy of bridges and on discussions
with engineers from the RAP panel (59). These discussions
included previous experience with impact damage on structures that resulted in loss of load-carrying capacity for a
prestressed bridge member.
The feature under the bridge, i.e., if the bridge were over
a roadway, was included as a factor to consider based on the
perceived risk of affecting the feature under the bridge. For
example, if a primary bridge member lost load-carrying
capacity or deteriorated to a serious condition, consequences
may be increased either as a result of falling debris or significant displacement, or emergency shoring that may be required
that would affect the serviceability of the roadway below
the bridge.
Additional factors considered for determining the CF
included considering the traffic volumes; in these analyses,
bridge decks with ADT greater than 10,000 were considered
to have High CFs. This is intended to reflect a case where deck
damage resulted in a major serviceability consequence.

The environmental conditions considered in the analysis of bridges in Oregon differed according to the damage
mode being considered. For example, for corrosion of super
structure metals (rebar or strands), the RAP identified three
separate areas with coastal and mountainous regions being
the most aggressive environment, while desert portions of the
state represented the least aggressive environment, obviously.
However, for spalling of bridge decks, the panel identified
areas of the state where de-icing chemical use was highest
because these areas are urban areas with high traffic volumes.
For the damage mode of rutting, travel corridors that experience high traffic volumes likely to be using studded tires were
identified. Generally, these corridors were identified because
they connected major urban areas and resort locations. The
environments identified by the Oregon RAP are summarized
in Table 13.

3.7.2CFs
There were six different CF cases considered in analyzing
results in Oregon, as shown in Table 14. These different cases
were selected to illustrate how different criteria established
by an RAP might affect the outcome of the analysis. These
included considering all superstructure damage modes as

Table 13. Environments identified by the Oregon RAP for different


damage modes.
Damage
Mode
Corrosion

Spalling

Rutting

Environment
Coastal and Mountainous
Valley or General
Environment
Desert
Portland
Salem
Bend
La Grande
I-5

I-84

Reason
Aggressive environment, high humidity and/or use of deicing chemicals

High application of de-icing chemicals

Presence of traveling traffic with studded tires

177
Table 14. CF cases used for back-casting
in Oregon.
Case No.
1
2
3
4
5
6

Description
High consequence for superstructure damage modes
Moderate consequence for superstructure damage
modes
Superstructure damage mode CF is determined by
redundancy and facility under bridge (screening not
used)
Superstructure damage mode CF is determined by
redundancy and facility under bridge screening
for CS 4 or 5 is used
All criteria in scenario 3 plus deck damage has high
consequence if ADT > 10000, screening not used
All criteria in scenario 5 plus considering screening
factors for CS 4 or 5

High consequence, considering all superstructure damage modes as Moderate consequence, and determining the
CF based on the redundancy of the bridge, as described in
Section 3.6.9. Additional analysis was done to test the effect
of including, or not including, the screening criteria of elements with a condition state identified as CS 4 or 5. It should

be noted that including this screening factor affects the OF,


making the likelihood High for any element with any portion of the element reported in CS 4 or 5, a failed condition.
Using these screening criteria does not change the CF, but
may change the inspection interval. This case, which includes
redundancy, feature under, and condition screenings is applicable for the subject bridges, and is shown in bold in Table 14.
Finally, the CF was adjusted to consider the consequences
for deck damage modes as High for bridges with high ADT,
in this case determined by bridges with ADT of 10,000 or
greater (according to NBI data). This case demonstrates the
consideration of traffic volumes in the consequence analysis
of a deck, which may be applicable in certain urban areas.

3.7.3 Back-Casting Results for Oregon


Figure 19 illustrates the results of the back-casting procedure as done on one of the subject bridges. Shown in this figure is the NBI condition rating history for the bridge, showing
how the condition ratings have varied over the course of the
back-casting period. On this graph, the inspection interval

Figure 19. Example of the back-casting process showing NBI condition ratings
over time and the inspection interval determined through RBI analysis.

178

determined through the RBI analysis for each year there was
an available element-level inspection report is shown enclosed
in a box near the bottom of the figure. In a few isolated cases,
there were not element-level reports available for every year,
though NBI data was available. This example was selected as
an illustration of applying the RBI analysis for each historical
inspection result, and how that outcome may vary over the
course of the life of a bridge. In this case, the inspection interval was reduced and then later increases following a repair,
based on the condition of the bridge. This was not common
occurrence, but it is useful as an illustration of how the results
of the back-casting are summarized in the figure, and how the
RBI inspection interval could vary over the life of the bridge
based on the RBI analysis. Also shown on the graph are any
repairs that had been completed on the bridge, and the year
that these repairs were completed.
It is very important to recognize that the RBI process is
not intended to predict or track the NBI ratings. In some cases,
changes in the inspection interval determined from the RBI
analysis may coincide with changes in the NBI condition rating,
because either these ratings are included in the analysis or the
rating changes coincide with changes in the element condition states that are included in the analysis. In other cases,
these may not coincide, because the RBI analysis depends not
only on the current condition, but also the potential for serious damage to occur looking forward based on the bridge
attributes (as expressed through the OFs) and the consequence of that damage. For example, a bridge rated in good
condition according to the NBI condition rating may have a
relatively short inspection interval, either because the potential for damage is high based on the attributes of the bridge,
or the consequences are high based on the redundancy or other
circumstances influencing the CF. The research team believes
this feature, i.e., the ability to look forward with an RBI analysis, is a significant advantage over the present calendar-based
system. At present, in the current calendar-based approach
there is no rational way to attempt to address the negative (or
positive) attributes associated with future condition of a given
specific bridge or family of bridges.
Overall, the results of back-casting verified that the methodology was capable of determining an effective and safe
inspection interval. There were no instances of bridge deteriorating to a serious condition during the RBI inspection
intervals recommended using the proposed methodology.
The process was effective in differentiating inspection intervals based on the risk profiles developed through the RAP
process, i.e., the OFs stemming from attribute scoring and
the CFs. In some cases, bridges that were in generally good
condition according to the NBI ratings resulted in short
inspection intervals, indicating that the process was sensitive
to risk factors that are not necessarily revealed through condition ratings. In other words, even though the condition of

the bridge at the present time was generally good, there was
a high likelihood of deterioration based on the design, environment, or loading of the bridge. In other cases, bridges that
included components rated in fair condition were assigned
longer intervals.
Table 15 shows the overall results for each of the CF cases
for the last inspection record analyzed for the 22 bridges, typically from an inspection conducted sometime between 2011
and 2013. The CF Case 4 is highlighted in the table because
this case, which includes consideration of the redundancy of
the bridge, traffic under the bridge, and screening any bridges
with elements with CS 4 or 5 reported, is a durable and widely
applicable category. These data are based on the consequence
cases described above and the data models developed through
the RAP. The year of construction, superstructure type (simple span or continuous), the facility under the bridge, and
the scour rating are also shown in the table. These data were
obtained from the NBI data for these bridges. This table also
presents results for Cases 1 and 2, with CF for the superstructure always high or always moderate, respectively. These data
represent the simplest analysis of the CF for a superstructure.
Cases 5 and 6, which included an ADT criteria for deck CF are
also shown, to illustrate how a more restrictive criteria for the
deck would affect the analysis.
Scour ratings were not a part of the RBI analysis, as scour
generally has its own evaluation procedures. Additionally, the
scour rating was not considered in the overall analysis because
this is a unique characteristic of the specific bridge, and therefore may skew the results for a population of bridges selected
at random. A bridge owner may choose to screen bridges with
poor scour ratings as a policy; however, screening bridges in
this manner in the current analysis would not be beneficial
in measuring the overall effectiveness of the RBI procedures.
Table 16 shows the summary of the RBI results for the
Oregon bridges in terms of percentage of the sample population. Based on these analyses, again focusing on CF Case 4,
approximately 41% of bridges would remain on a biennial
inspection schedule, while just over 59% of bridges would
have a larger interval of 48 or 72 months. These data illustrate
the effect of using different criteria to identify the CF for the
population of bridges, and results were as expected: relatively
simple but conservative use of CF of high for the superstructure results in fewer bridges identified with extended
intervals, using a less conservative moderate factor results
in more bridges on extended intervals, etc.
The overall results of the back-casting, considering each
of the analyses conducted at each existing inspection record,
are shown in Table 17. These results include 157 separate
analyses done based on the inspection records, and for each
of the six cases for determining the CF and OF described in
Table 14 above. CF Cases 5 and 6, which include consideration
of the ADT on the bridge deck, show only a modest differ-

179
Table 15. Overall results for each of the CF cases in Oregon.

1973
1973
1996
1987
1987

9546

1967

00988A
01056A
9358
16873
18175

1967
1970
1965
1991
1999

01895A

1995

9915
8994
8896
20666

1970
1962
1963
2009

19739

2007

19738

2006

19284

2005

Case 6

1962

Case 5

01741B
07935A
07935B
17451
16454
16453

Water
Highway
Relief for
waterway
Water
Water
Water
Highway
Highway
Highway/
waterway
Water
Water
Highway
Water
Water
Railroad
waterway
Highway
Water
Water
Water
Railroad
waterway
Railroad
waterway
Other

Scour
Rating

Case 4

1975
1973

Simple
span
(SS) or
Cont.
(C)
SS
C

Case 3

02376B
07801A

Facility
Under

Case 2

Year
Built

Case 1

Bridge
ID

3
N

48
24

48
24

48
24

48
24

24
24

24
24

SS

24

24

24

24

24

24

SS
SS
C
SS
SS

3
3
8
N
N

24
24
48
24
48

48
48
48
48
48

48
48
48
24
48

48
48
24
24
48

48
48
48
24
48

48
48
24
24
48

24

48

24

24

24

24

C
C
SS
SS
C

5
5
N
8
8

24
24
24
48
48

48
48
48
72
48

48
48
24
72
48

48
24
24
72
48

48
24
24
48
48

48
24
24
48
48

24

48

48

48

48

48

C
SS
SS
SS

N
U
3
8

24
24
48
48

48
48
48
72

24
24
48
72

24
24
48
72

24
24
48
48

24
24
48
48

24

48

48

48

24

24

24

48

48

48

24

24

48

48

48

48

48

48

Note: N = not over waterway, U = bridge with unknown foundation.

ence. The results shown in this table are generally consistent


with those shown in Table 16, considering that the bridges
are aging with time, and consequently the inspection intervals
may be reduced. For example, at the end of the back-casting
period, 50% of the bridges had a 48-month inspection interval
assigned, as shown in Table 16. However, 68% of the bridges
had a 48-month interval assigned at some point in the backcasting period, and 57% of all of the analyses conducted indicated a 48-month interval, as shown in Table 17.

Table 16. Summary of final back-casting


intervals for 22 bridges in Oregon.
CF Case No.
1
2
3
4
5
6

24 month
64%
9%
32%
41%
45%
55%

Inspection Interval
48 month
72 month
36%
0%
82%
9%
59%
9%
50%
9%
55%
0%
45%
0%

These data are significant in showing the consistency of the


process when applied over 17 years of historical data through
the back-casting process.
Significantly, there were no instances of bridge deteriorating to a serious condition between inspection intervals, and
those with poor condition rating generally were assigned
inspection intervals of 24 months based on the RBI analysis.
For example, Figure 20 presents the condition rating history
and RBI inspection interval for Bridge 16454. This bridge was
Table 17. Summary of back-casting
intervals for 22 bridges in Oregon
(all analyses).
CF Case No.
1
2
3
4
5
6

Inspection Interval
24 month
48 month
72 month
62%
38%
0%
8%
82%
10%
28%
62%
10%
34%
57%
9%
39%
58%
3%
44%
53%
3%

180

Figure 20. Condition rating history and RBI inspection interval


for Bridge 16454.

constructed in 1987, less than 30 years ago, and the back-casting


assessment for the bridge was initiated in 1998, when the bridge
was only 11 years old. However, the RBI inspection interval was
determined to be 24 months, due to damage modes related to
corrosion susceptibility of the superstructure. For this bridge,
cracking in the superstructure was present early in the service
life, resulting in increased likelihood of corrosion damage to
the strands in the prestressed members. A repair completed
in 2007 consisted of epoxy injection of the superstructure
cracking. Looking forward from 1998, the superstructure
condition deteriorated relatively rapidly as the bridge aged.
For this bridge, the RBI assessed interval was 24 months
throughout the back-casting period, an appropriate interval
given the susceptibility to corrosion damage for this bridge.
The validity of the short interval is also supported by the fact

that the CR decreased from 6 to 4 around 2002. Again, the


ability of the RBI method to identify the attributes that would
suggest the superstructure is susceptible to damage resulted
in the shortened interval.
There were several bridges that had reported poor condition ratings, and typically those had inspection intervals of
24 months assigned. There were some exceptions: for example, Bridge 07935A (Figure 21) had a reported condition rating of poor (CR = 4) in 2013 and had an overlay installed,
and the inspection interval assigned by RBI was 48 months.
This may seem like a long inspection interval considering this
deck apparently required an overlay. However, the elementlevel condition state for the deck was 100% in CS 2 (CS 2 =
Patched areas and/or spalls/delaminations exist on either side
of the deck. The combined distressed area is 10% or less of the

181

Figure 21. NBI condition rating history and RBI inspection intervals
for Bridge 07935A.

total deck area); the soffit element was 95% in CS 1 and 5%


in CS2, and the deck cracking element was 100% in CS 1. In
this case, the assigned NBI condition rating appears not to be
well correlated with the element condition states. Given that
the NBI condition rating typically have a variability of 1,
and in this case was not consistent with the element-level
data, it may be that the condition rating is not reflective of the
overall conditions. Since these element-level condition states
contribute significantly to the likelihood estimate, a longer
inspection interval was assigned.
Arrows superimposed on the figure illustrate when an RBI
inspection would have been conducted, assuming the start year
of 1999. In this case, the year of the RBI inspection would not
coincide with the year that the condition rating of 4 occurred,
though the schedule year is somewhat arbitrary, being based
herein on the earliest date of available data. This example was
the most problematic of the 22 sample bridges included in
the back-casting, in terms of the RBI interval assigned for the
bridge. However, as described above, the apparent incongruity

between the RBI inspection interval and the condition rating


was explained by the element-level inspection results.
3.7.3.1 Risk Matrices
The results of the analysis can be illustrated on the risk
matrix to summarize the data and indicate the controlling damage modes, i.e., those damage modes representing
the highest risk or IPN. Table 18 shows the damage modes
assessed in the Oregon case study, along with an alpha-numeric
Table 18. Key to risk matrix summaries of RAP analysis.
Deck
Spalling (D1)
Rutting (D2)
Cracking (D3)

Superstructure
Cracking (S1)
Strand Corrosion (S2)
Impact (S3)
Rebar Corrosion Within the
Span (S4)
Bearing Seat Problems (S5)

Substructure
Settlement (F1)
Corrosion (F2)

182

Figure 22. Risk matrix for Bridge 16454 illustrating


results of the RBI analysis.

identifier (D1, D2, etc.). Figure 22 includes the risk matrix for
Bridge 16454 with the damage modes located on the matrix
according to the results of the RBI analysis (OF and CF). As
shown in this figure, the results of the RBI analysis are plotted in appropriate locations on the diagram. The locations on
these plots describe the inspection interval identified, and can
also be used to calculate the IPN to identify the most important damage modes as identified through the RBI process. For
example, in the plot shown, the IPN for S1, S2, and S4 = 9,
indicating that these damage modes (cracking, strand corrosion, and rebar corrosion) have high importance related to
the risk profile for the bridge. These data are useful for identifying emphasis areas for the inspection of the bridge, and
could be included in inspection procedures or guidance as
a normal outcome of the RBI assessment. Such risk-based
inspection procedures may improve the reliability of inspection and communicate the engineering-based RBI assessment
of the key damage modes for a bridge to inspectors in the field.
Appendix C includes the controlling damage modes for the
RBI analysis of bridges in Oregon. Frequently, several of the
damage modes had similar risk profile, such that there is not a
controlling damage mode. This is typical for bridges in good
condition, such that inspection intervals are typically longer.
These controlling damage modes evolve during the service life
of the bridge, as damage develops and affects the OF.
3.7.3.2 Surrogate Data for a Family of Bridges
An analysis was conducted of the overall inventory in
Oregon based on the results of the RAP analysis. The objec-

tive of this analysis was to identify the population of low-risk


bridges that were in very good condition and that could be
assessed in an entirely data-driven process that did not require
individual assessments of a bridge. Such bridges could beconsidered for extended inspection intervals throughout the RBI
analysis, based only on a screening process that utilized data
in existing databases. These included a series of 22items that
were readily available, such as NBI items or bridge elements
included in standard inspection reports. Table 19 indicates
the individual items that were analyzed and the accepted
values from the screening. Each of the criteria was based on
attributes or items developed from the RAP analysis. Each
of these items is shown in Table 19, along with the screening
criteria used to analyze the inventory data. Generally, these
criteria include bridges that have NBI condition ratings of 7
or higher and have no elements with any condition states of
3 or higher reported. In this case, scour ratings were considered as shown in the table, eliminating bridges with unknown
foundations, bridges without scour analysis, or bridges that
are scour critical.
Screening the Oregon databases was performed by the
Oregon DOT, which provided a listing of all bridges meeting
the element-level screening criteria included in the table. For
the NBI criteria, filtering of the data was performed by the
research team. Generally, these database searches and filtering took only a short time intervala matter of 1 hour or
less, consisting primarily of inputting screening or filtering
criteria in search functions and yielding immediate results.
The results of the analysis indicated that 18% (652/~3600)
of the prestressed bridge inventory met all of the criteria indicated in the table. For these bridges, the likelihood of serious
damage developing in the next 72 months interval could be
considered low or even remote, based on the RAP analysis.
Assuming the CF to be moderate for this population of bridges,
an inspection interval of 72 months could be assigned. If the
effect of scour is not considered, or considered as a separate
inspection requirement, the number of bridges meeting the
other criteria was 970 bridges, or about 1 in 4 bridges.
These data indicate that the RAP process can be used to
develop criteria for an entirely data-driven process for identifying bridges that are very low risk, and the number of bridges
meeting these criteria is significant (almost 1 in 5 prestressed
bridges in Oregon). Such analysis takes only a matter of a few
hours to complete, once the data items are identified through
the RAP process.

3.8 Back-Casting Results for Texas


This section of the report describes the results of backcasting for steel bridges in the state of Texas. This includes
a description of the environments identified by the RAP for
use in the OF analysis, the CF used in the back-casting analy-

183
Table 19. List of criteria for data-driven screening process based on RBI.

No.
1
2
3
4
5
6
7

Prestressed Bridges (5, 6)*


Deck
Item
Criteria
Damage Mode
#358 Deck Cracking SF
CS 2 or less
Deck Cracking
#359 Soffit Cracking SF
CS 2 or less
Spalling
Deck Elements
CS 2 or less
Spalling
Less than 50
Age
Spalling
years
NBI Item 58
7 or greater
Spalling
# 325
CS 2 or less
Spalling
Coded 1 or
#370-374
Fire
uncoded

#326

CS 1 only

Rutting

Notes

Deck Condition
Deck Condition
Dynamic Loading
Fire or Incident
Deck Wearing
Surface Condition

Superstructure

Superstructure
Impact
Cracking

Legal Load Capacity

Impact

No Overtopping

NBI Item 54

17 ft. or greater

10

NBI Item 70

11

NBI Item 71

Coded 5
Coded 4 or
greater

12

NBI Item 41

Coded A

Cracking

13

#362 Impact(SF)

None

Rebar Corrosion

14

Superstructure Elements
#104, 109, 115

CS 2 or less

15

NBI Item 59

7 or greater

16

Deck Joint Items (All)

CS 2 or less

17

Bearing Elements (All)

CS 2 or less

18

NBI Item 34

19

#360 Settlement SF

20

NBI Item 60

7 or greater

21

NBI Item 113

Not U, 6 or 0-4

Settlement

22

Substructure Elements

CS 2 or less

Corrosion Damage

30 degrees or
less
Substructure
CS 1 or
uncoded

Bridge Height

Open, No
Restrictions
Traffic Impact
Smart Flag

Strand and Rebar


Corrosion, Bearing
Superstructure
Condition
Bearing Area
Damage
Bearing Area
Damage
Bearing Area
Damage

Failed Deck Joint

Settlement

Settlement

Corrosion Damage

Superstructure
Condition Rating

Bearing Issues
Bridge Skew

Substructure
Condition
Scour
Sub. Element
Conditions

5 = prestressed concrete, 6 = continuous prestressed concrete: from NBI database.

sis, overall results, and specific examples selected to illustrate


implementation of the technology.

3.8.1Environments
The environmental conditions considered in the analysis of
bridges in Texas also differed depending on the damage mode
being considered for the RAP. Generally, the RAP identified
an east-west interstate highway, I-20, as dividing the state into
areas where de-icing chemical were likely to be used (north)
and areas where they are very unlikely to be used (south).
These environments were applied for most damage modes,
such as spalling of bridge decks. For the damage mode of section loss in steel members, the RAP identified that the coastal
areas were the most aggressive environment, followed by areas
north of I-20 and a moderately aggressive environment, and
all other areas being the least aggressive environment.

3.8.2CFs
There were four different CF cases considered in analyzing
results in Texas, as shown in Table 20. These different cases
were selected to illustrate how different criteria established by a
RAP might affect the outcome of the analysis. These included
considering all superstructure damage modes as high
Table 20. CF cases used for back-casting in Oregon.
Case No.
1
2
3
4

Description
High CF for superstructure
Superstructure CF is determined by redundancy and
facility under bridge (screening not used)
Superstructure damage mode CF is determined by
redundancy and facility under bridgescreening
for pin and hanger used
All criteria in scenario 3 plus deck damage has high
consequence if ADT > 10000

184

Case 1

Case 2

Case 3

Case 4

Year Built

Table 21. List of bridges analyzed in Texas.

01-139-0-0769-01-007
02-127-0-0014-03-194
02-127-0-0094-04-057
02-220-0-1068-02-058
05-152-0-0067-11-188
08-030-0-AA01-31-001
12-085-0-1911-01-003
12-102-0-0027-13-195

1956
Waterway
1963
Highway
1939
Waterway
1957
Highway
1990 Highway, Railroad
1985
Waterway
1943
Waterway
1979
Highway

C
C
SS
C
C
SS
SS
SS

5
N
8
N
N
5
8
N

24
24
48
24
48
48
24
48

24
24
72
24
48
48
48
48

24
24
72
24
48
48
48
48

24
24
48
24
48
48
48
48

12-102-0-0500-03-320
15-015-0-0025-02-162
15-015-0-B064-55-001
18-057-0-0092-14-210
18-061-0-0196-01-133
19-019-0-0610-06-162
23-141-0-0251-05-020
23-215-0-0011-07-056
24-072-0-0167-01-059

1990
1967
1964
1973
1960
1971
1934
1948
1970

C
C
C
C
C
C
C
C
C

N
N
5
N
N
N
8
8
N

48
48
48
48
24
24
48
48
48

48
48
72
48
24
24
48
48
48

48
48
72
24
24
24
48
24
48

24
48
72
24
24
24
48
24
48

Bridge ID

Facility Under

Highway
Highway
Waterway
No Feature Under
Highway
Highway
Waterway
Waterway
Highway

Structure
Type

Scour
Condition

consequence, and determining the CF based on the redundancy of the bridge, as described in Section 3.6.9. Additional
analysis was done to test the effect of including, or not including, the screening criteria for a bridge with a pin and hanger
connection. This screening criteria were not identified during
the RAP process, although it would likely have been identified
during the course of a full-scale implementation of RBI. Again,
this screening factor affects the OF, making the likelihood high
for any component containing a problematic detail such as a
pin and hanger. Using these screening criteria does not change
the CF, but may change the inspection interval. Finally, the CF
was adjusted to consider the consequences for deck damage
modes as high for bridges with high ADT, again determined by
bridges with ADT of 10,000 or greater (according to NBI data).
The CF Case 3 is considered the most appropriate case for the
analysis, and is highlighted in the following tables.

had a 72-month inspection interval and 53% with a 48-month


inspection interval, while 35% were found to have a 24-month
maximum interval. The maximum interval found during the
analysis indicated that 24% of the bridges had an RBI interval of
72 months at some point during the back-casting period, indicating that the RBI practice included shorter intervals as these
bridges became older and deterioration progressed.
Appendix C includes the controlling damage modes for
the RBI analysis of bridges in Texas. Frequently, several of the
damage modes had similar risk profile, such that there is not a
controlling damage mode. This is typical for bridges in good
condition, such that inspection intervals are typically longer.
Table 23 shows the overall results from each of the 117 analyses
conducted during the back-casting procedure. These data illustrate the relative consistency of the process and the application
of the attribute criteria to the steel bridge population in Texas.

3.8.3 Back-Casting Results for Texas

3.8.3.1Examples

Table 21 lists the bridges analyzed in this portion of the


study. This table includes data on the year of construction, the
facility under the bridge, the structure span type (simple or
continuous), etc. The CF Case 3 is highlighted in this table to
illustrate the most likely or commonly applicable CF case that
would be utilized to evaluate the bridges.
Data models developed through the RAP process were used
to analyze each bridge and determine the appropriate RBI interval. Table 22 shows the results of the analysis for themost recent
year for which inspection results were available. As shown in
the table, for the most recent analysis year, 12% of the bridges

This section provides two examples from the analysis of


bridges in Texas. The first example is a bridge that included
Table 22. Results of back-casting for bridges
in Texas.
CF Case No.
1
2
3
4

Inspection Interval
24 month
48 month
72 month
35%
65%
0%
29%
59%
12%
35%
53%
12%
47%
47%
6%

185
Table 23. Results of back-casting including
all analysis.
CF Case No.
1
2
3
4

Inspection Interval
24 month
48 month
72 month
29%
71%
0%
27%
58%
15%
32%
53%
15%
44%
50%
6%

a pin and hanger connection. In the back-casting analysis,


the presence of a pin and hanger connection was used as a
screening factor that made the OF high, regardless of other
attributes of the bridge. This screening factor is based on the
historical experience that pin and hanger connections frequently present maintenance challenges. Figure 23 indicates
the inspection intervals determined for the structure during
the back-casting, along with the NBI condition rating history.
As can be seen in the figure, the superstructure condition
rating dropped 3 ratings, from 7 to 4, over a single inspection
interval. According to the inspection records reviewed during
the back-casting, this reduction was due to damage to the pin
and hanger connection. Rehabilitation of this pin and hanger
joint was required and was ongoing in 2013.

This example is important for illustrating the importance


of identifying screening factors in the RBI process. Screening
factors are intended to identifying bridge attributes that make
the likelihood of serious damage unusually high, unusually
uncertain, or otherwise different than other bridges in a group.
As shown in this example, the screening factor of bridges with
pin and hanger connections was needed to capture the unusual
behavior of this bridge.
The second example was a steel multi-girder short-span
bridge constructed in 1943. For this bridge, located in a coastal
environment, back-casting indicated an inspection interval of
72 months between the years of 2001 and 2006, changing to a
48-month interval based on the results of the 2008 inspection.
The change in the inspection interval for this bridge resulted
from corrosion-related deterioration of the superstructure,
i.e., likelihood for severe section loss. As shown in Figure 24,
even though this bridge was 70 years old, the overall condition of the superstructure was satisfactory at the beginning
of the back-casting period, and subsequently reduced to fair,
where the structure condition rating remained. The inspection interval is also reduced during this period. Again, the RBI
practice does not necessarily reflect the NBI condition ratings,
as the data model includes specific information regarding the

Figure 23. Historical NBI data and RBI inspection intervals for a
steel bridge in Texas with a pin and hanger connection.

186

Figure 24. Example bridge in Texas with decreasing


inspection interval resulting from section loss.

element condition state and other factors. It should also be


noted that the deck and substructure are generally in satisfactory condition, and the superstructure is in Fair condition
due to the damage mode of section loss caused by corrosion,
known to be a slow-acting deterioration mechanism.
Table 24 indicates the damage modes evaluated for the steel
bridges in Texas, including damage modes for the superstructure, deck, and substructure. Figure 25 indicates these damage
modes plotted on the standard risk matrix, with Figure25A
being the risk matrix for the bridge including a pin and
hanger connection, and Figure 25B the bridge with section
loss. Considering Figure 25A, the data plotted on the figure

illustrate that according to the damage modes identified, the


inspection interval for this bridge would be 48 months. Recall
that this bridge included a pin and hanger connection, used
as a screening criteria to identify the OF as high for the superstructure. In other words, the data in Figure 25A indicates the
inspection interval for the bridge if the bridge did not include
a pin and hanger connection. This illustrates how screening
criteria affect the analysis; for this bridge, the overall condition based on the condition rating, notes, and element-level
data suggest an inspection interval of 48 months. However,
the bridge includes an attribute, i.e., a pin and hanger connection, that makes the anticipated behavior of the bridge

Table 24. Damage modes for the steel bridges in Texas.


Deck
Spalling (D1)
Punch Through (D2)
Cracking (D3)
Delamination (D4)

Superstructure
Section Loss (S1)
Impact (S2)
Fatigue Cracking (S3)
Overload Damages (S4)

Substructure
Settlement (F1)
Corrosion (F2)

187

Figure 25. Risk matrices for steel bridges in Texas.

unusually uncertain, and not typical of other bridges in the


family. As such, this screening criterion is critical to determining the effective interval for this bridge.
Figure 25B indicates the risk matrix for Bridge 1922-01-003.
As shown in this figure, the damage mode of section loss controls the inspection interval for the bridge. These data illustrate
how individual damage modes can control the inspection
interval for the bridge. In this case, the bridge is 70years old,
and in a relatively aggressive coastal environment. As such, it
is rational that a shorter inspection interval would be required
than if the bridge were in an arid environment.

3.8.3.2 Surrogate Data


Surrogate data was analyzed for Texas based on the data
models developed from the RAP. Because Texas has not traditionally used its element-level data for bridge management purposes, and as such these data are not maintained within a single
database, the surrogate data relied solely on NBI data to scan
the inventory and identify bridges for the extended interval of
72 months. Table 25 indicates the parameters used in the analysis.
Based on this analysis, it was found that 927 bridges, or
12.5% of the inventory in Texas, met all of these parameters

Table 25. List of criteria for data-driven screening process based on RBI
for Texas.

No.

Item

Steel Bridges (3, 4)


Deck
Criteria
Damage Mode
Less than 50
Spalling
years
7 or greater
Spalling, Cracking
Superstructure

Notes

Age

NBI Item 58

10

NBI Item 70

Coded 5

Cracking

11

NBI Item 71

Coded 4 or
greater

Legal Load
Capacity

Impact

No Overtopping

12

NBI Item 41

Coded A

Cracking

Open, No
Restrictions

NBI item 54

17 ft or greater

15

NBI Item 59

7 or greater
30 degrees or
less
Substructure

Superstructure
Impact
Superstructure
Condition
Bearing Area
Damage

18

NBI Item 34

20

NBI Item 60

7 or greater

Corrosion Damage

21

NBI Item 113

Not U, 6 or 0-4

Settlement

Deck condition
Deck Condition

Bridge Height
Superstructure
Condition Rating
Bridge Skew
Substructure
Condition
Scour

188

for the extended inspection interval. These data were also analyzed without regard to the age criteria identified in Table25.
This resulted in a slight increase: 1068/7423 = 14.38%.

3.9Discussion of the Case Studies


in Texas and Oregon
The case studies were used to verify the effectiveness of the
RBI procedure developed through the research. Overall, the
back-casting illustrated that the RBI process was effective in
determining a suitable inspection interval for each bridge in
the study.

3.9.1 Back-Casting Results


The back-casting procedure was used to verify the effectiveness of the RBI process, and there were three primary questions addressed, as discussed in Section 3.6.4. The following
discusses each question individually in terms of the outcome
of the back-casting.
1. Did the condition rating for any component change significantly during the RBI interval in a manner that was not captured or anticipated effectively, but would have been captured
(or detected sooner) by a standard, 24-month interval?
A detailed review of the condition ratings for each of
the bridges included in the study was conducted, as illustrated in the examples presented herein. This review and
analysis indicated that there were no cases where the condition rating changed unexpectedly in a manner that was
not captured or reflected in the RBI inspection interval
identified when screening criteria were used. Recall that
screening criteria of CS 4 or 5 for prestressed bridges in
Oregon, and a pin and hanger connection in Texas, were
implemented in the analysis.
2. Were there any significant maintenance or repair actions completed that would have been delayed as a result of implementing
an RBI interval (relative to a standard, 24-month interval)?
Reviews of the repair histories for the subject population
of bridges were conducted based on available records. This
review did not indicate any instances where there were sudden or unexpected repairs required that would have been
delayed as a result of RBI intervals. There were cases where
routine maintenance or repair, such as a deck overlay, may
not coincide with an RBI interval; however, this depends on
when the RBI cycle was initiated. There were also several
cases where repair or rehabilitation activities were performed

during the back-casting window; however, the activities were


generally consistent with the RBI analysis, and would not be
adversely affected by the RBI implementation. For example,
a bridge identified by RBI as being susceptible to corrosion
damage had epoxy injection performed, consistent with the
RBI analysis. In most cases, there were no significant repairs
during the back-casting window.
3. Were there any significant factors or criteria not identified
through the RAP analysis that were needed in the data models
to provide suitable results?
There was one case in each state where there were factors that were not identified through the RAP processes
were needed for the data models. In Oregon, a screen for
elements with CS 4 or 5 was needed in the data models;
in Texas, a screen for pin and hanger connections was
needed. In both cases, these are relatively obvious additions to the data models that were overlooked during the
RAP meetings, but would likely be identified by anyone
implementing the back-casting procedures.
The overall objectives of the case studies were to demonstrate
the implementation of the methodologies with state DOT personnel, and verify the effectiveness of RBI analysis in determining suitable inspection intervals for typical highway bridges.
In terms of these objectives, the RAP meeting in each state,
and the effectiveness of the data models developed through
that RAP process, indicated that the processes developed for
RBI analysis were effective, practical, and implementable using
state DOT personnel in Texas and Oregon. The results of the
back-casting process described above verified the effectiveness
of the RBI procedures, and demonstrated that implementation of the RBI practice did not adversely affect the safety and
serviceability of the sample bridges analyzed.
It should also be noted that for bridges where an inspection
interval of 72 months was proposed using the RBI analysis,
there were no cases of sudden repair, unexpected progression
of damage, or sudden changes to the condition ratings for the
bridge. These results indicated that the RBI procedures were
effective in identifying a portion of the inventory, typically on
the order of 10% of the sample bridge population, where an
inspection interval of 72 months provided a suitable inspection interval that did not compromise the safety and serviceability of these bridges. It should be noted that the sample
population of bridges was older than the average age of the
inventories in each state, such that the identified rate (~10%)
would likely be higher for a population of bridges constructed
more recently.

189

CHAPTER 4

Conclusions, Recommendations,
and Suggested Research
This research developed inspection practices to meet the
goals of (1) improving the safety and reliability of bridges and
(2) optimizing resources for bridge inspection. The goals of
the research have been achieved through the development
of a new guideline document entitled Proposed Guideline
for Realibility-Based Bridge Inspection Practices, which has
been developed based on the application of reliability theories. This document meets the project objective of developing a recommended practice for consideration for adoption
by AASHTO, which is based on rational methods to ensure
bridge safety, serviceability, and effective use of resources.
A reliability-based approach was fully developed and documented through the Guideline. Background information and
foundation for key elements of the process have been further
expanded in the present report, to provide additional details
and perspectives on the research conducted as part of the project. However, the primary outcome of the study is the comprehensive Guideline developed, which provides a new paradigm
for bridge inspection. This new paradigm could transform
the calendar-based, uniform inspection strategies currently
implemented for bridge inspection to a new, reliability-based
approach that will better allocate inspection resources and
improve the safety and reliability of bridges.
The implementation of the Guideline developed through the
research was tested by conducting case studies in two states.
The objectives of the case studies were to demonstrate the
implementation of the methodologies with state DOT personnel, and verify the effectiveness of RBI analysis in determining
suitable inspection intervals for typical highway bridges. The
verification of the methodology was analyzed using a backcasting procedure that compared historical inspection records
and the results of RBI analysis. These studies demonstrated
and verified the effectiveness of the procedures developed in
the research for identifying appropriate inspection intervals for
typical highway bridges. It was shown through these studies
that the RBI practices identified appropriate inspection intervals of up to 72 months. It was concluded from these studies

that implementation of the RBI practices did not adversely


affect the safety and serviceability of the bridges analyzed in
the study, based on the analysis of historical inspection records.
These studies also successfully demonstrated the implementation of the Guideline and the procedures therein using state
DOT personnel.
The results reported herein demonstrated and verified that
inspection intervals of up to 72 months were suitable for certain
bridges. Such extended inspection intervals would allow the
reallocation of inspection resources toward bridges requiring
more frequent and in-depth inspections, resulting in improved
safety and reliability of bridges. As such, the project goals of
developing a reliability-based bridge inspection practice that
could improve the safety and reliability of bridges, and optimizes the use of resources, were achieved through the research.
The following sections describe specific recommendations
and suggested research, including detailed suggestions for conducting key elements of an implementation strategy intended
to support the broader implementation of the research.

4.1Recommendations
The research reported herein has demonstrated the effectiveness of the RBI procedures for determining suitable
inspection intervals for typical highway bridges, and as such,
implementation of the RBI technology is recommended. The
research also demonstrated that inspection intervals of up to
72 months were suitable for certain bridges and did not affect
the safety and serviceability of bridges analyzed in the study.
Such extended inspection intervals would allow the reallocation of inspection resources toward bridges requiring more
frequent and/or in-depth inspections, resulting in improved
safety and reliability of bridges. Based on these results, implementation of RBI technology and inspection intervals of up
to 72 months for certain bridges should be pursued.
The procedure, methods, and approach described herein
can be applied for atypical bridges as well. For example,

190

non-redundant bridge members can be assessed using this


approach, as illustrated in previous research (60). Specific
attributes may differ for such an application; examples and
illustrations of applying the RBI technology for these applications should be pursued. The approach can also be applied to
complex bridges, or to bridges with advanced deterioration.
Analysis requirements may be more detailed and advanced;
development of such analysis should be pursued to provide
a uniform strategy for bridge inspection across the entire
bridge inventory. Additional research and testing may be used
to broaden the application of the RBI technology.
Finally, the back-casting procedure utilized herein should
be considered for implementation when RBI practices are
to be used. This recommendation is based on the research
result that indicated a screening criteria in each state was
not identified during the RAP. Additionally, the RAP process
may be subject to variability not observed in the research
when applied over a broader platform. Back-casting provides
a means for verification of models developed by the RAP and
QA tool for assessing the RBI process. As such, the backcasting procedure provides a critical tool for the implementation of RBI technology.

4.2 Suggested Research


Suggested research stemming from this project includes
developing applications of the technology for atypical highway
bridges, including non-redundant members, complex bridges,
and bridges with advanced deterioration, as described above.
Additional research to demonstrate the consistency of the
process across a larger population of bridge owners, and for
families of bridges not examined herein, should be undertaken.
In addition, efforts will be required to support implementation of the technology. A comprehensive implementation
plan, which includes additional research on such factors as
economics of applying the methodology, is included below.

4.2.1 Implementation Strategy


The implementation of a reliability-based inspection planning process such as described herein will be a difficult challenge.
As with any existing established procedures, specifications, or
policy, change is difficult. The current U.S. bridge inspection
program and associated procedures have been the standard
since the early 1970s, and considerable infrastructure has
been developed to support the program. Well-established
training, experience, and organizational structure within
state departments of transportation will need to be modified to meet the needs of an RBI practice. The workforce will
need to be retrained to meet the needs of the new approach
to inspection and inspection planning. Existing legislative
requirements, including the NBIS will need to be modified to

allow for the new methodology to be implemented. Therefore, a strategy for converting the established bridge inspection programs from a uniform, calendar-based system to a
reliability-based system is required. This section of the report
describes implementation strategies and tasks to establish a
new paradigm for bridge inspection based on the RBI processes described in the Guideline.
A number of implementation challenges exist looking forward toward the adoption of the RBI methodology. Inspection
program organizational structures and personnel may need
to be modified to accommodate the larger role of engineering and inspection planning required for RBI compared to
a uniform, calendar-based approach. Personnel with suitable
experience and knowledge to effectively conduct the necessary
assessment will be required. In an era where government agencies are suffering significant fiscal challenges, often resulting
in staff reductions, developing and retaining the necessary
resources may be a challenge. A strong technical foundation
for RBI will need to be developed to justify maintenance of
the resources needed.
Training and knowledge development to support RBI will
also be needed to implement the technology on a widespread
basis. Developing the necessary tools to train individuals in
the various aspects of the technology and processes will be an
important part of technology transfer and implementation.
There will also be a significant political challenge to modifying an existing inspection system, which has been in effect for
many years, with a process that may result in fewer inspections
for certain bridges, even if the process results in an improvement in the safety of bridges overall. Engineers, inspectors.
and maintenance personnel are likely to perceive the benefits of a more rational system, but the non-technical audience may be more difficult to convince. Data from additional
case studies or pilot implementation, economic impacts, and
safety analysis will be required to provide evidence to support
the new approach for inspection planning. However, because
deterioration patterns for bridges typically require a long time
period to manifest, and failures are rare, generating empirical
data to measure improvements in safety will be a significant
challenge. The implementation strategy described in this section is designed to address these issues, and will require some
investment of resources to execute and complete effectively.
Given these challenges, the implementation strategy has
been developed to meet the following goals:
Provide a technical foundation for widespread implemen-

tation of the technology and


Develop community support for the new inspection

approach.
Activities in the implementation strategy to provide a
technical foundation for widespread implementation of the

191

technology include conducting additional case studies in certain states to test and develop the technology further, developing training modules and software to support the technologies,
and conducting a study focused on the economic and safety
impacts of transitioning to the new inspection approach.
To develop community support, the implementation plan
proposes developing an oversight committee to monitor and
develop the Guideline and address bridge-owner needs, and
developing an effective communication strategy. Throughout
the implementation activities described herein, the FHWA
can play an important role in assisting with moving the technology toward eventual acceptance.
4.2.1.1 Implementation Tasks
The strategy developed for implementing RBI for bridges
in the United States will require a number of steps be completed to test and refine the technology, develop support for
transition to a new approach to inspection planning, and
eventually gain widespread acceptance of the new technology.
The implementation plan developed for the project consists
of a number of individual tasks to be completed to achieve
the desired goals.
Task 1. Establishment of an oversight committee.
Task 2. Additional case studies.
Task 3. Development of training modules.
Task 4. Develop a communications strategy.
Task 5. Economic and safety impact study.
Task 6. Software development and integration
The sections that follow address each of the implementation tasks to be completed toward widespread adoption
of the RBI technology.
4.2.1.2Task 1. Establishment of
an Oversight Committee
An important element of the longer-term implementation of
RBI practices will be the establishment of a committee structure
to oversee the development and maintenance of the technology.
Implementation of the proposed methodology will require a
significant shift in paradigm for inspection planning for highway bridges. Consequently, there will need to be a long-term
commitment with respect to maintaining and implementing
the new methodologies contained in the RBI Guideline. As is
common with many design codes and standards, a committee
is needed to oversee, maintain, and further develop the Guideline. This committee may be a subcommittee of the AASHTOs
Standing Committee on Bridges, or a subcommittee of the
existing T-18 committee on Bridge Management, Evaluation
and Rehabilitation, or even the committee itself.

The committee should have the goal of providing objective oversight and management of the Guideline and requirements for RBI. Committee membership should be diversified,
and include representatives from states in different geographic
regions and with different types of bridge inventories. Participation of the FHWA in this committee would be desirable.
During the transitional stages, which should be anticipated,
the committee should include both states implementing or
developing RBI processes, and states that are not yet utilizing RBI. Care should be taken to ensure the participation of
the community as a whole in the committee, including both
bridge owners adopting the RBI Guidelines and those who
have not yet made the transition. An important aspect of the
proposed methodology is transparency, and any committee
overseeing progress will require critical voices to be effective.
The role of the committee should be as follows:
1. Oversee implementation of the technology across different
states and act as a focal point for information interchange
regarding states experience, research, and developments.
2. Identify and recommend research and development needs
to support the technology.
3. Recommend and approve changes to the RBI Guideline
document.
It is envisioned that the oversight committee or sub
committee will be a long-term or even permanent organization that will serve the larger bridge community.
4.2.1.3 Task 2. Additional Case Studies
Additional case studies may be needed to test the application of the Guideline, identify implementation challenges, and
provide additional data on the impact of transitioning to an
RBI approach. The objectives of the case study should include:
Assess the effect of RBI outcomes on the inspection prac


tice for different families of bridges.


Evaluate implementation challenges.
Assess the repeatability and consistency of the process.
Provide baseline data for economic and safety impact study.
Revise the RBI guideline as needed.

The focus of these case studies will be to evaluate processes


and methods described in the Guideline. To meet the objectives shown above, case studies should be conducted in several
states to evaluate different families of common bridge types.
4.2.1.4 Task 3. Development of Training Modules
Implementation of the RBI process will obviously require
the development of training modules for those that will be

192

involved in the process. The inspection planning process is


more involved and complex under an RBI scheme relative
to a calendar-based inspection planning process. The assessment of reliability characteristics requires an understanding
of the approach and the assessment needs. Therefore, training for both members of the RAP and for inspectors that will
implement the results of the RBI planning process may be
necessary. This section provides an overview of the training
needs for implementing RBI practices.
4.2.1.4.1 Training for RAP Member. The development
of an expert panel like the RAP is relatively new to the industry, and will require a substantial commitment from stakeholders to identify and train individuals to participate in the
process. Individuals that can provide expertise objectively are
needed, and training in the tools and mechanism for providing
that objective expertise will be required to ensure the methodology is effectively implemented. Although these panels must
be objective, they should also collectively possess an intimate
knowledge of the inventory of the given agency or DOT. This
group must also be isolated from the political and management pressures that could undermine the objectivity and effectiveness of the process (e.g., pressure to use the RAP process
to simply extend intervals to save money).
Effective training for RAP members will be one of the most
critical parts of the implementation of RBI on a broad scale.
This training should provide sufficient knowledge in the theory
and underlying approach to RBI planning, deterioration and
reliability science, methodologies for expert elicitation, and
processes for determining the factors required for the analysis.
Training modules developed during the case studies that were
conducted as part of the research were shown to be effective,
based on the results of the case studies, and they provide a
strong foundation for the development of more formal training for widespread implementation of the technology.
4.2.1.4.2 Training of Inspectors. Implementation of
RBI practices will require training for bridge inspectors to
develop the necessary understanding of the RBI process. RBI
assessments for inspection planning provide a prioritization
of inspection needs for a bridge based on the anticipated or
expected damage modes and the importance of that damage
in terms of safety of the bridge. Criteria developed through
the RAP process identify key condition attributes used to
determine the reliability of individual elements of the bridge
and related criteria for reassessment of the inspection inter
val and scope. Additionally, the IPN identified through the RAP
analysis prioritizes damage modes for inspection in a manner that is significantly different than the traditional, detect
all the defects approach. It is critical that the underlying
approach and methodologies used in the planning process

are understood by the inspectors implementing the practices


to ensure adequate inspection in the field to support the overall process. Training of this type was not addressed during the
course of the research.
Significant resources for inspector training already exist,
and are generally implemented through the National Highway Institute (NHI). Existing training modules will need to
be adjusted to accommodate the focus on damage and damage precursors that are a part of the RBI process. Inspector
training modules like the 2-week inspector training course
and supporting Bridge Inspectors Reference Manual will need
to be modified to be more focused to incorporate the perspective of the RBI process and its approach to ensuring the
safety and reliability of bridges.
Entirely new training modules could be developed to support the RBI Guideline. However, while the approach to inspection planning and the required reporting and inspection results
differ for RBI, the damage modes that affect bridges are typically well-covered in the existing training modules. Developing
entirely new training for RBI including all of the information
and examples already included in the existing modules would
be a duplication of effort that is likely unnecessary. However,
there are certain subjects not currently included in the existing training modules that will be required to effectively implement RBI. Table 26 describes two training modules that may
be necessary for implementing RBI. This training for inspectors describes the underlying concepts and methodologies for
RBI. It is intended that the training modules be presented
at the appropriate technical level to develop sufficient background knowledge for an inspector that will implement the
RBI process. Ultimately, these specific modules may be added
to the existing curriculum to provide training continuity and
avoid duplication.
The training modules included in Table 26 are intended to
include enhanced training in specific inspection methodologies required for implementing RBI, including those identified in the RBI Guideline. For example, increased training
for visual inspection to detect fatigue cracking, appropriate
lighting and distance requirements, and thoroughness of
inspection should be addressed through the training. Implementation of the other basic techniques, such as sounding or
concrete, should also be included in the training.
More advanced technologies, such as advanced NDE techniques that may be specified for certain damage modes, will
also require training. For example, if the RAP identifies the
use of infrared thermography as a means of assessing delaminations in concrete, then specialized training in applying this
technology will be needed. Training in advanced NDE technologies is typically advanced and focused, and utilization
of the technology is specialized in nature. Specific training
in these technologies should be developed as needed to meet
specific owner needs.

193
Table 26. Outline of training for inspectors.
Module I Background
Topics
Deterioration mechanism for
bridges
Fundamentals of reliability
theory and application to
inspection
Reliability assessments for
RBI

Notes
Overview of typical deterioration patterns
Background overview of the underlying theories for RBI,
reliability matrices, and likelihood
RAP process and basis for inspection procedures
Module II Practices

Understanding the IPN


Inspection needs, criteria, and
reporting
Enhanced inspection methods
for RBI

Required thoroughness of inspection and prioritization of damage


modes
Focus and scope of inspections for RBI, access requirements,
reassessment criteria, documentation, and reporting requirements
Technologies and methods for detecting identified damage
modes, enhanced methods for RBI, sounding and crack detection

4.2.1.5 Task 4. Develop a Communications Strategy


As discussed, the proposed method is a significant change
in paradigm for the bridge inspection community, and as
such developing an effective communications strategy will be
a key element of overall success. Education of policy makers
and DOT administrators as to the benefits of the proposed
methodologies will be needed. Although it is anticipated that
owners (i.e., state bridge engineers) will likely embrace the
proposed methodologies, it will require the buy-in of policy
makers to actually implement any changes to the bridge
inspection program. This group includes DOT administrators as well as appointed or elected officials. Since the current
inspection program is covered by the Code of Federal Regulations (CFRs), lack of approval of policy makers could restrict
any proposed methods from being implemented. Constant
communication with the FHWA regarding the methodology
and its progress will be very important to providing decision
makers the information necessary to support future changes
that may be needed. Additional interactions with state and
local government rules, and potential conflicts with other
specifications will also need to be assessed.
There also exists the challenge that policy makers may have
difficulty separating gross numbers of inspections from quality and effectiveness of inspection. Numbers of inspections are
frequently equated to safety, even though these two factors
may be unrelated, particularly when the method of inspection is ineffective for a particular damage mode or deterioration mechanism. There will be a need for clear explanation of
the approach to achieve buy-in from the policy makers, and
even then challenges should be expected.
There also exist the potential that the cost reallocations
will be misinterpreted as a reduction in inspection requirements to save money, rather than a reallocation of resources
to be most effective in ensuring bridge safety. If viewed as
a cost saving measure, the practice could lead to reduction in

available resources for inspection, which is undesirable. Care


needs to be taken to illustrate the enhanced reliability realized
through allocating resources more effectively, and the benefits in terms of supporting the inspection and repair needs of
an aging bridge inventory. Public and political acceptance of a
system that may result in fewer inspections will rely on clearly
communicating the benefits (more in-depth and focused
inspections), not any cost saving.
The communication strategy developed should include the
development of non-technical publications that describe the
RBI approach and highlight the benefits such as increased
resources to focus where most needed and reductions in the
risks and costs associated with unnecessary inspections. The
improved reliability and safety of bridges that can be realized
through improved inspection practices should be described
as well as the improved management and responsible utilization of public funds highlighted in these publications.
Technology transfer to the broader engineering community should also be developed as part of the communications
strategy.
4.2.1.6Task 5. Assessment of the Economic
and Safety Impacts of RBI
A key element in pursuing widespread implementation
and acceptance of the RBI technology will be a critical assessment of the economic and safety impacts of converting from
a uniform, calendar-based system to the RBI methodology.
Such a study will likely be a required component of gaining
the support of AASHTO and policy makers, who would naturally question the cost and safety impact of such a transition.
Because inspection resources are reallocated and optimized under the RBI process, an organized and systematic
assessment of the effect of the process on bridge safety will
be required. This study should examine both the benefits
of increased inspection thoroughness and assess any real or

194

potential diminishment of safety or safety effects on a given


bridge inventory associated with varying inspection intervals to match the needs for bridges. The study should also
examine the safety effects of continuing the current status
quo, addressing both the cost and safety implications of the
do-nothing approach, including the effects of decreasing
available resources for inspection. Data from the case studies
can be used to assist in this assessment process for this study.
Study of the economic impact of transitioning to RBI is
also needed. Because the methodology requires the investment of increased resources for the planning of inspections,
bridge owners may need to restructure traditional responsibilities and staffing to address the needs of full implementation of the technology. Increased engineering efforts are
required to complete RAP analysis, particularly in contrast
to uniform, calendar-based approaches. Inspections under
RBI Guidelines typically have increased scope and increased
access requirements relative to traditional routine inspections, and as such are likely to have increased costs. On the
other hand, the inspection may be less frequent, such that the
overall costs may be unaffected. The economic implications
for transitioning to RBI obviously will vary according to the
current inspection practice currently used in a state, and
additional information on the actual or estimated economic
impacts will be needed.
This study of the economic and safety impacts of transitioning to RBI practices will likely be a key tool to the eventual
political and policy acceptance of the new technology. This
implementation activity is high priority as a means of addressing the issues associated with achieving acceptance of the

new technology among the public, with policymakers, and


with stakeholders.
4.2.1.7Task 6. Software Development
andIntegration
An important element of widespread implementation and
acceptance of the new technology will be the development of
software tailored to meet the needs of the process, and intended
to integrate current or future data collection and storage
approaches used by bridge owners. The processes for assessing the OFs, such as identifying and scoring key attributes of
bridges, can be repetitive once established, and therefore lend
themselves to software implementations. Many of the attributes
identified through the analysis process may already be stored in
existing databases and bridge management systems. Condition
attributes and screening criteria may be implemented through
existing software developed for bridge inspection and storing
bridge inspection data, or in new software developed with RBI
in mind. Such software is widespread in other industries and
used for risk assessment and condition-based maintenance of
facilities and components. The process of implementing a RBI
practice can be simplified by the development of software to
more rapidly utilize the methodology. Integration with existing software and databases that store relevant information
will be beneficial for efficiency in implementation. The case
studies conducted as part of the research reported herein
developed some basic software tools for these purposes; these
tools will need to be integrated into existing software. Developing software to assist in the RBI process will be necessary
for implementation efforts to be successful.

195

References

1. National Bridge Inspection Standards, in 23 CFR part 650. 2004:


USA. p. 7441974439.
2. FHWA, Revisions to the National Bridge Inspection Standards (NBIS).
1988: p. 21.
3. Washer, G., Connor, R., Ciolko, A., Kogler, R., Fish, P., and Forsyth, D.,
Developing Reliability Based Inspection Practices: Interim Report.
2009, NCHRP: Washington, D.C. p. 76.
4. AASHTO, The Manual For Bridge Evaluation. 2008, AASHTO Publications: Washington, D.C.
5. Estes, A.C., D.M. Frangopol, and S.D. Foltz, Updating Reliability
of Steel Miter Gates on Locks and Dams Using Visual Inspection
Results. Engineering Structures, 2004(26.3), Elsevier: p. 319333.
6. Estes, A.C. and D.M. Frangopol, Updating Bridge Reliability Based
on Bridge Management Systems Visual Inspection Results. Journal
of Bridge Engineering, 2003. 8(6), ASCE: p. 374382.
7. Estes, A.C. and D.M. Frangopol, Bridge Lifetime System Reliability under Multiple Limit States. Journal of Bridge Engineering, 2001.
6(6), ASCE: p. 523528.
8. Sommer, A.M., A.S. Nowak, and P. Thoft-Christensen, Probability-Based Bridge Inspection Strategy. Journal of Structural Engineering, 1993. 119(12), ASCE: p. 35203536.
9. Stewart, M.G. and J.A. Mullard, Spatial time-dependent reliability
analysis of corrosion damage and the timing of first repair for RC
structures. Engineering Structures, 2007. 29(7), Elsevier: p. 14571464.
10. Darmawan, M.S. and M.G. Stewart, Spatial time-dependent reliability analysis of corroding pretensioned prestressed concrete
bridge girders. Structural Safety, 2007. 29(1), Elsevier: p. 1631.
11. Stewart, M.G., D.V. Rosowsky, and D.V. Val, Reliability-based
bridge assessment using risk-ranking decision analysis. Structural
Safety, 2001. 23(4), Elsevier: p. 397405.
12. Chung, H.Y., Fatigue Reliability and Optimal Inspection Strategies for
Steel Bridges in Civil Engineering. 2004, University of Texas: Austin,
Texas.
13. Enright, M.P. and D.M. Frangopol, Service-Life Prediction of
Deteriorating Concrete Bridges. Journal of Structural Engineering,
1998. 124(3), ASCE: p. 309317.
14. Estes, A.C. and D.M. Frangopol, Repair Optimization of Highway
Bridges Using System Reliability Approach. Journal of Structural
Engineering, 1999. 125(7), ASCE: p. 766775.
15. Akgl, F. and D.M. Frangopol, Time-dependent interaction between
load rating and reliability of deteriorating bridges. Engineering Structures, 2004. 26(12), Elsevier: p. 17511765.
16. Straub, D. and M.H. Faber, System Effects in Generic Risk
Based Inspection Planning. ASME Conference Proceedings, 2002.
2002(36126): p. 391399.

17. Goyet, J., A. Rouhan, and M.H. Faber, Industrial Implementation


of Risk Based Inspection Planning Lessons Learnt From Experience: Part 1The Case of FPSOs. ASME Conference Proceedings,
2004. 2004(37440): p. 553563.
18. Poulassichidis, T. Risk-Based Inspection as a Reliability-Engineering Tool for Fixed Equipment Decisions, in 2004 ASME Pressure
Vessel and Piping Conference. 2004. San Diego, California: Dow
Chemical.
19. Faber, M.H., et al., Field Implementation of RBI for Jacket Structures. Journal of Offshore Mechanics and Arctic Engineering, 2005.
127(3), ASME: p. 220226.
20. Frangopol, D.M., J.S. Kong, and E.S. Gharaibeh, Reliability-Based
Life-Cycle Management of Highway Bridges. Journal of Computing
in Civil Engineering, 2001. 15(1), ASCE: p. 2734.
21. ASME, Inspection Planning Using Risk-Based Methods. 2007, The
American Society of Mechanical Engineers. p. 92.
22. Regulatory Guide 1.178An Approach to Plant-Specific Riskinformed Decisionmaking for Inservice Inspection of Piping, N.R.
Commission, Editor. 2003: Washington, D.C. p. 15.
23. API, API Recommended Practice 580, Risk-Based Inspection. 2002
American Petroleum Institute: Washington, D.C. p. 45.
24. Bowles, D.S., et al., Portfolio Risk Assessment: A Tool for Dam Safety
Risk Management, in 1998 USCOLD Annual Lecture. 1998, Utah
State University: Buffalo, New York.
25. FHWA, Recording and Coding Guide for the Structural Inventory
and Appraisal of the Nations Bridges, U.S.DOT, Editor. 1995, U.S.
Department of Transportation.
26. AASHTO, AASHTO Bridge Element Inspection Manual. 2010,
AASHTO Publications: Washington, D.C. p. 170.
27. Cambridge Systematics, I., PONTIS 4.4 Users Manual. 2004,
AASHTO: Washington, D.C.
28. Stewart, M.G. and D.V. Rosowsky, Structural Safety and Serviceability of Concrete Bridges Subject to Corrosion. Journal of Infrastructure Systems, 1998. 4(4), Elsevier: p. 10.
29. API, Risk-Based Inspection Technology, API Recommended Practice
581. 2008.
30. Straub, D., et al., Benefits of Risk Based Inspection Planning for Offshore Structures. ASME Conference Proceedings, 2006. 2006(47489):
p. 5968.
31. ABS, Surveys Using Risk-Based Inspection for the Offshore Industry.
2003, American Bureau of Shipping: Houston, TX.
32. Faber, M.H., et al., Fatigue Analysis and Risk Based Inspection
Planning for Life Extension of Fixed Offshore Platforms. ASME
Conference Proceedings, 2005. 2005(41952): p. 511519.

196
33. Straub, D. and M.H. Faber, Computational Aspects of Risk-Based
Inspection Planning. Computer-Aided Civil and Infrastructure
Engineering, 2006, Wiley: p. 179192.
34. Moore, M., et al., Reliability of Visual Inspection for Highway Bridges.
2001, Federal Highway Administration: McLean, VA.
35. Agrawal, A., A. Kawaguchi, and C. Zheng, Bridge Element Deterioration
Rates, TIRC/NYSDOT, Editor. 2009, The City College of New York,
Department of Civil Engineering: Albany, New York. p. 105.
36. Ehlen, M.A., M.D.A. Thomas, and E.C. Bentz, Life-365 Service
Life Prediction Model Version 2.0.1 Users Manual. 2009, Life-365
Consortium II.
37. Steel Bridge Design Handbook, Chapter 23, Corrosion Protection of
Steel Bridges. 2010, National Steel Bridge Alliance.
38. Albrecht, P. and J.T.T. Hall, Atmospheric Corrosion Resistance of
Structural Steels. Journal of Materials in Civil Engineering, 2003.
15(1), ASCE: p. 224.
39. Vu, K.A.T. and M.G. Stewart, Structural reliability of concrete
bridges including improved chloride-induced corrosion models.
Structural Safety, 2000. 22(4), Elsevier: p. 313333.
40. Sun, X., et al., Analysis of Past National Bridge Inventory Ratings
for Predicting Bridge System Preservation Needs. In Transportation
Research Record: Journal of the Transportation Research Board, No.
1866, Transportation Research Board of the National Academies.
2004. p. 3643.
41. Russell, H.G., NCHRP Synthesis 333: Concrete Bridge Deck Performance. 2004, Transportation Research Board of the National Academies, Washington, D.C. p. 188.
42. Ramey, G.E. and R.L. Wright, Bridge Deterioration Rates and
Durability/Longevity Performance. Practice Periodical on Structural Design and Construction, 1997. 2(3): p. 98104.
43. A. Sohanghpurwala, NCHRP Report 558: Manual on Service Life of
Corrosion-Damaged Reinforced Concrete Bridge Superstructure Elements. 2006, Transportation Research Board of the National Academies, Washington, D.C.
44. Albrecht, P. and J. Terry T. Hall, Atmospheric Corrosion Resistance
of Structural Steels. Journal of Materials in Civil Engineering, 2003.
15(1), ASCE: p. 224.
45. Klinesmith, D.E., R.H. McCuen, and P. Albrecht, Effect of Environmental Conditions on Corrosion Rates. Journal of Materials in
Civil Engineering, 2007. 19(2), ASCE: p. 121129.

46. Administration, NASA, Nondestructive Evaluation Requirements for Fracture-critical Metallic Components. 2008, NASA:
Washington, D.C.
47. Boring, R.L. A Review of Expertise and Judgement Processes for Risk
Estimation, in European Safety and Reliability Conference. 2007.
Stavanger, Norway.
48. Boring, R., Gertman, D, Joe, J., Marble, J., Galyean, W., Blackwood, L.,
Blackman, H., Simplified Expert Elicitation Guideline for Risk Assessment of Operating Events. 2005, Nuclear Regulatory Commission:
Washington, D.C. p. 16.
49. API, Risk-Based Inspection, API Recommended Practice 580. 2002:
p. 45.
50. Andersen, G.R., et al., Risk Indexing Tool to Assist in Prioritizing
Improvements to Embankment Dam Inventories. Journal of Geotechnical and Geoenvironmental Engineering, 2001. 127(4), ASCE:
p. 325334.
51. Ayyub, B.M. Uncertainties in Expert-Opinion Elicitation for Risk
Studies. 2000. Santa Barbara, California, USA: ASCE.
52. Hetes, B., et al., Expert Elicitation Task Force White Paper. 2009, EPA
Science Policy Council: Washington, D.C.
53. Bulusu, S. and K. Sinha, Comparison of Methodologies to Predict
Bridge Deterioration. Transportation Research Record: Journal of the
Transportation Research Board, No. 1597, 1997, p. 3442.
54. Hong, T.-H., et al., Service life estimation of concrete bridge
decks. KSCE Journal of Civil Engineering, 2006. 10(4), Springer:
p. 233241.
55. ASTM, Corrosiveness of Various Atmospheric Test Sites As Measured
by Specimens of Steel and Zinc, in Metal Corrosion in the Atmosphere, W.H.A.S.K. Coburn, Editor. 1968. p. 397.
56. FHWA, Characterization of the Environment. 2000, Washington, D.C.
57. FHWA, Characterization of the EnvironmentRevisit of Exposure
Sites in the Continental US. 2003. Washington, D.C.
58. Moore, M.E., Phares, B.M., Graybeal, B.A., Rolander, D.D., and
Washer, G.A., Reliability of Visual Inspection of Highway Bridges,
FHWA, Editor. 2001, U.S.DOT: Washington, D.C.
59. Ghosn, M.M., Redundancy in highway bridge superstructures. 1998,
Washington, D.C.
60. Connor, R.J. and M.J. Parr, A Method for Determining the Interval for
Hands-On Inspection of Steel Bridges with Fracture Critical Members.
2008, Purdue University. p. 32.

197

Abbreviations
ADT
Average Daily Traffic
ADTT
Average Daily Truck Traffic
BME
Bridge Management Element
BMS
Bridge Management Software
CF
Consequence Factor
CFR
Code of Federal Regulation
CIF
Constraint-Induced Fracture
CS
Condition State
CVN
Charpy V-Notch
DOT
Department of Transportation
EMAT
Electromagnetic-Acoustic Transducer
GPR
Ground Penetrating Radar
HPC
High Performance Concrete
HPS
High Performance Steel
HS
High Strength
IE
Impact Echo
IPN
Inspection Priority Number
IR
Infrared Thermography
LFD
Load Factor Design
LIBS
Laser-Induced Breakdown Spectroscopy
MT
Magnetic Particle Testing
NBE
National Bridge Elements
NBI
National Bridge Inventory
NBIS
National Bridge Inspection Standards
NDE
Nondestructive Evaluation
NHI
National Highway Institute
OF
Occurrence Factor
PCI
Precast/Prestressed Concrete Institute
PDF
Probability Density Function
POD
Probability of Detection
POF
Probability of Failure
PT
Dye Penetrant Testing
QA
Quality Assurance
QC
Quality Control
RAP
Reliability Assessment Panel
RBI
Risk-Based Inspection
SIP Stay-in-Place
SI&A
Structural Inventory and Appraisal
TRL
Technical Readiness Level
UPV
Ultrasonic Pulse Velocity
UT-T
Ultrasonic Thickness Gauge

198

APPENDIX A

Developing Reliability-Based Inspection


Practices: Oregon Pre-Stressed Bridges
199 Bridge/Deck/Spalling
199 Bridge/Deck/Rutting
200

Bridge/Deck/Cracking (Non-Corrosion Induced)

200

Bridge/Superstructure/Cracking (Shear)

201

Bridge/Superstructure/Strand Corrosion

201

Bridge/Superstructure/Fire Damage

202 Bridge/Superstructure/Impact
202

Bridge/Superstructure/Rebar Corrosion within the Span

202

Bridge/Superstructure/Bearing Seat Problem(s)

203 Bridge/Substructure/Settlement
203

Bridge/Substructure/Corrosion Damages (Spalling/Delamination/Cracking/Rust)

199

C.9 and C.12

C.10 and C.11

Cracking/Spalling

High

Medium

Low

Remote

Condition

#358 CS 4
Or
#359 CS 4 or
CS 5

#358 CS 3
Or
#359 CS 3

#358 CS 2
Or
#359 CS 2

#358 CS 1
Or
#359 CS 1

Delamination/Patch

Condition

Screening
Degree of
Severity
Max
Score

Similar Items in
Guideline

Type of
Attributes

Attributes

Bridge/Deck/Spalling

15

Source of data

#358 Deck Cracking Smart Flag


#359 Soffit Cracking Smart Flag
(Oregon Coding Guide Pages 79
and 80)

20

<500

15

Concrete Decks and Slabs without


an Overlay : #12 - #26 -#27 -#38 #52 -#53
Concrete Decks or Slabs with a
Thin or Rigid Overlay: #18 - #22#44 - #48.
Item 29 NBI

Desert

20

Bridge File

<10 years

20

Item 27 NBI (Year Built)

20

# 325 (Oregon Coding Guide


Page 22)

>25%
CS 4 or CS 5

11%-24%
CS 3

<10%
CS 2

501-4999
Valley
(general
environment)
10-49 years
<40mph +
CS 2 or CS 3
+ <40mph

CS 1

L.1

ADTT

Loading

>5000

L.3 (Exposure
Environment)

Location
/Environment

Loading

Coastal and
Mountain

D.6 (Year Built)

Age

Design

>50 years

L.2

Dynamic Loading

Loading

>40 mph +CS 3

C.21 and C.13

Rebar Corrosion

Condition

Rust/Black/Low
Cover

No stains,
Epoxy/high
cover

20

Concrete Elements(Oregon
Coding Guide Page 38-41)

L.5

De-icing

Loading

High (Regions
like Portland,
Bend, Salem,
La Grand )

Low (All
Other
Regions)

15

Items 3, 4, and 5 NBI, or


Geographical map

CS 2 +
<40mph

CS 1

ADT

Loading

>15000
vpd

D.10

Wearing
Surface Type

Design

AC

I-5 highway
Portland to
Salem and I-84
Portland

L.3
(potential to be
exposed to high ADT
with studded tires)

C.2

Location
Current
Condition
(amount of
rutting)

Loading

Condition

Low

100014999 vpd
Bare
Concrete/S
TR
overlay/Ep
oxy
All other
locations

Present
(>0.5")

Remote

<1000vpd

Max
Score

L.1

Medium

Degree of
Severity

Type of
Attributes

High

Screening

Similar Items in
Guideline

Attributes

Bridge/Deck/Rutting
Source of data

20

Item 29 NBI

Open Grid

15

Item 108A NBI


(Also page 120
Oregon Coding
Guide)

20

Items 3, 4, and 5
NBI, or
Geographical map

None
(<0.5")

4H
&
4M
(M)
+

15

(Oregon Coding
Guide Pages 22 &
23)

200

Medium

Low

Remote

Unsealed
cracks exist
in the deck
that are of
moderate
size (0.025
to 0.060 in.
wide) and
density (3
to 10
apart).

Unsealed cracks
exist in the deck
that are of
moderate size
(0.025 to 0.060
in. wide) or
density (3 to 10
apart).

The surface of the


deck is cracked,
but the cracks are
either filled/sealed
or insignificant in
size and density to
warrant repair
activities.

C.9

Cracking

Condition

Unsealed cracks
exist in the deck
that are of severe
size (>0.060 in.
wide) and/or
density (<3
apart)

D.18
L.1
D.20

Skew
ADTT
Thickness

Design
Loading
Design

>30 o
>5000
<7"

501-4999

<30o
<500
>7"

L.2

Profile/
Dynamic
Loading

Loading

>40 mph +
CS 3

<40mph +
CS 2 or
CS 3 +
<40mph

CS 2
+ <40mph

S.10

Span Type

Screening

CS 1

Max
Score

High

Degree of
Severity

Screening

Type of
Attributes

Similar Items
in Guideline

Attributes

Bridge/Deck/Cracking (Non-Corrosion Induced)


Source of data

20

#358 Deck
Cracking Smart
Flag (Oregon
Coding Guide Page
79)

M
H
H

15
20
20

Item 34 NBI
Item 109 NBI
Bridge File

20

Item 325 (Oregon


Coding Guide Page
22)

Continuous or Non Continuous

Loading

D.18
D.6 (Year of
Construction)

Skew

Design

Age

Design

D.20

AASHTO
Shear
Design

Screening

AASHTO
requirements were
not considered in
design

Remote

Other

Max
Score

Overload

Low

Degree of
Severity

L.4 and D.2

If it has already
posted for less than
legal load or exposed
to overload

Mediu
m

Screening

Type of
Attributes

High

Similar Items in
Guideline

Attributes

Bridge/Superstructure/Cracking (Shear)
Source of data

20

Item 41 NBI
(See also Oregon Coding
Guide on page 95 )

>30

<30

10

Item 34 NBI

<2000

>2000

10

Item 27 NBI (Year Built)

AASHTO
requirements
were
considered
in design

201

C.8 (Corrosion
Induced Cracking)
C.1

High

Medium

Low

ENV

Loading

Coastal
and
Mountain

Valley
(general
environment)

Desert

Existing
Damage

Condition

CS 4

CS 3

CS 2

Current
Condition

Condition

5 and less

6
between 1.5"
and 2.5"

D.11 (Minimum
Concrete Cover)

Cover

Design

1.5" or
Less,
Unknown

D.12 (Reinforcement
Type)

Strand Type

Design

Uncoated

D.20 and S.10

Bad End Detail

Design

Has Strand
Exposure
to outside
environment

Remote

Bridge/Superstructure/Fire Damage
Reason(s) for Attribute
Incidences of fire on or below a highway bridge are not
uncommon. This type of damage is most frequently caused
by vehicular accidents that result in fire, but secondary causes
such as vandalism, terrorism, or other damage initiators
should not be discounted. If fire does occur on or below a
bridge, an appropriate follow-up assessment should be conducted to determine how the fire has affected the load carrying capacity and the durability characteristics of the main
structural members and the deck. This assessment is typically
performed during a damage inspection immediately following the incident.
Damage to bridge components resulting from a fire is
either immediately apparent during the damage inspection,

Source of data

20

Geographical Map

20

Prestressed/Post Tensioned
Concrete Elements
(Oregon Coding Guide Page 40)

7 or
greater

20

Greater
than or
equal 2.5"

20

Bridge File

10

Bridge File

10

Bridge File

CS 1

Do not
have
Exposure
to outside
environment

Item 59 NBI
(See also page 42 and 104
Oregon Coding Guide)

Epoxy
coated

Unknown

Screening
Degree of
Severity
Max
Score

L.3 (Exposure
Environment)

Type of
Attributes

Similar Items in
Guideline

Attributes

Bridge/Superstructure/Strand Corrosion

or may manifest within the first 12-to-24 month interval following the fire. Based on this observation, bridges that have
experienced a fire may be screened from the reliability assessment until an inspection, which has been conducted approximately 12 months or more after the fire, confirms that the
fire has not affected the typical durability characteristics of
the bridge components. The purpose of this screening is to
ensure that damage from the fire has not manifested after the
damage inspection.

Assessment Procedure
This attribute is scored based only on the occurrence of a
fire on or below the structure being assessed. It is assumed
that an appropriate assessment immediately following the
fire incident (i.e., damage inspection) has been performed.

Fire incident has occurred and an inspection 12 months


after the fire has not occurred

Bridge is not eligible for reliability assessment until


inspection confirms that the bridge is undamaged

There have been no incidences of fire on or below the bridge,


or inspections conducted approximately 12 months or more
after the fire have confirmed that the bridge is undamaged

Continue with procedure

202

L.8

High Water

Low

Design

<15'

15-16

>17'

Remote

Max
Score

Medium

Degree of
Severity

Clearance

High

Screening

D.3

Type of
Attributes

Similar Items in
Guideline

Attributes

Bridge/Superstructure/Impact
Source of data

10

Item 10 NBI (Minimum


vertical under clearance)

Look at item 71
in NBI databaseif the code is 3 the
chance of over
top is occasional

Screening

Item 71 in NBI database


(See also page 117 Oregon
Coding Guide)

L.3 (Exposure
Environment)

ENV

Medium

Low

Loading

Coastal and
Mountain

Valley
(general
environme
nt)

Desert

#362 CS 2

C.6 and C.21


Previously Impacted
Active Corrosion

Existing
Damage

Condition

Prestressed
/Post
Prestressed/
Post Tensioned Tensioned
Concrete
Concrete
Elements
Elements

CS 4

CS 3

Between
1.5" and
2.5"

D.11 (Minimum
Concrete Cover)

Cover

Design

1.5" or Less,
Unknown

D.12 (Reinforcement
Type)

Strand Type

Design

Uncoated

Remote

Degree of
Severity
Max
Score

High

Screening

Similar Items in
Guideline

Type of
Attributes

Attributes

Bridge/Superstructure/Rebar Corrosion within the Span

#362 CS 1
Prestressed
/Post
Tensioned
Concrete
Elements

Prestressed
/Post
Tensioned
Concrete
Elements

20

Items 3, 4, and 5 NBI, or


Geographical map

#362-Traffic Impact
Smart Flag
(page 83 Oregon Coding
CS 3

Greater
than or
equal 2.5"
Epoxy
Coated

Guide)
Prestressed/Post Tensioned
Concrete Elements
(Oregon Coding Guide Page
40)

20

20

Bridge File or Cover


meter

20

Bridge File

CS 1

CS 2

Source of data

C.21

Corrosion

Condition

CS 4

D.18

Skew

Design

>30 o

C.22

Debris

Condition

Medium

CS 3

Flood region
Debris
INS.RPT

Max
Score

High

Screening
Degree of
Severity

Type of
Attributes

Similar Items in
Guideline

Attributes

Bridge/Superstructure/Bearing Seat Problem(s)


Source of data

H+

20

Prestressed/Post Tensioned
Concrete Elements
(Oregon Coding Guide
Page 40)

<30 o

10

Not
Susceptible

10

other

L+

10

Item 41 NBI
(See also Oregon Coding
Guide on page 95)

15

Bridge File

20

20

Low

CS 2

L.4

Overload

Loading

If it has already
posted for less
than legal load
or exposed to
overload

S.10

Design Details

Design

Simple Support

Continuous
Support

Integral
Abutments

C.4

Failed Joints

Condition

CS 3

CS 2

CS 1

C.2

Existing Damage

Condition

CS 3

CS 2

CS 1

Remote

CS 1

Jointless

Item 34 NBI
Item 113 NBI
(See also Oregon Coding
Guide on page 121)

Deck JointsOregon
Coding Guide Page 54-60
Bridge Bearing
ElementsOregon
Coding Guide Page 61-66

203

Medium

Low

D.21

Footing
Type

Design

Spread FTG on
soil/unknown
Foundation

Drill Shaft
friction Pile
/ETC

D.22

Subsurface
Condition

Condition

Slide zone, clay,


silt, shale, gravel

Limestone

C.3

Existing
Settlement

Condition

Active (No
monitor data)

S.10

Scour
Rating

Screening

4-6 (Oregon
Scour Code)

Remo
te

Degree of
Severity
Max
Score

High

Screening

Type of
Attributes

Similar
Items in
Guideline

Attributes

Bridge/Substructure/Settlement

If foundation was based on


Rock/Piles we do not need
to deal with other
following attributes

Source of data

20

Bridge File

solid, Rock

20

Bridge File

Occurred
but arrested

None

20

>7

<3

Item #360 on
page 81 Oregon
Coding Guide
Item 113 NBI
(See also
Oregon Coding
Guide on page
124)

C.8 (Corrosion
Induced Cracking)
C.1

High

Medium

Low

ENV

Loading

Coastal
and
Mountain

Valley
(general
environment)

Desert

Existing Damage

Condition

CS 4

CS 3

CS 2

Current Condition

Condition

5 and less

6
Between 1.5"
and 2.5"

D.11 (Minimum
Concrete Cover)

Cover

Design

1.5" or
Less,
Unknown

D.12 (Reinforcement
Type)

Rebar Type

Design

Uncoated

C.4

L.5

Failed Joints

De-icing

Condition

CS 3

Loading

High
(Regions
like
Portland,
Bend,
Salem, La
Grand)

CS 2

Remote

Screening
Degree of
Severity
Max
Score

L.3 (Exposure
Environment)

Type of
Attributes

Similar Items in
Guideline

Attributes

Bridge/Substructure/Corrosion Damages (Spalling/Delamination/Cracking/Rust)


Source of data

20

Geographical Map

20

Prestressed/Post Tensioned
Concrete Elements
(Oregon Coding Guide Page 40)

7 or
greater

20

Greater
than or
equal 2.5"

20

10

Bridge File

20

Deck JointsOregon Coding


Guide Page 54-60

15

Items 3, 4, and 5 NBI, or


Geographical map

CS 1

Low (All
Other
Regions)

CS 1

Item 59 NBI
(See also page 42 and 104
Oregon Coding Guide)

Epoxy
coated
Joint
less

Bridge File

204

APPENDIX B

Texas Steel Bridge Attributes Summary

205 Bridge/Deck/Spalling
205

Bridge/Deck/Punch Through

206 Bridge/Deck/Cracking
206 Bridge/Deck/Delamination
207 Bridge/Superstructures/Sectionless
207 Bridge/Superstructures/Impact
208

Bridge/Superstructures/Fatigue Cracking

208

Bridge/Superstructures/Fire Damage

209

Bridge/Superstructures/Deflection Overload

209

Bridge/Substructures/Corrosion Damages (Spalling/Delamination/Cracking/Rust)

209 Bridge/Substructures/Settlement

205

Low

Clear Cover

Design

<1"

1"-2"

>2"

20

Bridge file or Covermeter

Overlay

Design

Yes

No

10

Item 108A NBI (See also pages 9


and 10 Texas Coding Guide)

C.10

Delamination

Condition

Yes

No

20

Pages 5 and 8 Texas Coding


Guide

D.8

Mixed design
(Water)

Design

Poor Mix/Poor
H2O

All Else

15

Bridge file

ADTT
Thickness

Loading
Design

>5000
<7"

<5000
>8"

L
M

10
15

Item 29 & 109 NBI

D.20
D.19

Cold Joints

Design

Yes

No

15

Bridge file (or observation)

Condition

Yes

No

15

Pages 30 and 31 Texas Coding


Guide

Loading

Above I-20

All Else

20

Bridge file

Condition

50+

Other

15

Item 27 NBI (Year Built)

Screening

D.11

D.10

L.1

C.9
L.3
D.6

Cracking (map
dense)
Environment
Age Years of
Services

7"-8"

Degree of
Severity
Max
Score

Medium

Screening

Type of
Attributes

High

Similar Items in
Guideline

Remote

Attributes

Bridge/Deck/Spalling
Source of data

Bridge file

Type of
Attributes

High

Medium

Low

Degree of
Severity

Max
Score

Source of data

Thickness

Design

<7"

7"-8"

>8"

20

Bridge file

Map Cracking

Condition

Yes

No

20

Pages 30 Texas Coding


Guide (Deck Cracking)

C.10 and C.12

Delamination /
spall to rebar

Condition or
Screening if
more than
10%

Delamination
and spalling
>6%

Delamination
and spalling
<1%

15

Pages 5 and 8 Texas


Coding Guide

D.8

Poor Concrete Mix


(Poor Water)

Screening

15

Bridge file

L.1

ADTT

Loading

20

Item 29 NBI

10

Bridge file (PONTIS


Report)

Similar Items in
Guideline

D.20

C.9

L.3

Environment

Loading

Previous Punch
outs /rep

Screening/Yes
or No

>5000
Above I-20

Delamination
and spalling
2%-5%

<5000
All Else

Remote

Attributes

Bridge/Deck/Punch Through

Pages 5 and 8 Texas


Coding Guide

206

Existing Cracking

Condition

Yes

No

20

Page 30 Texas Coding Guide


(Deck Cracking)

D.20

Construction
Tech/Spec

Design

Bad

All Other

15

Bridge file

L.3

Environment

Loading

Above I-20

All Else

20

Bridge file

D.18 and D.19

Design Details
(Cold Joints, Skew)

Design

Yes

None

20

Bridge file

D.11

Cover

Design

<1"

1"-2"

>2"

20

Bridge file or covermeter

Medium

Low

1"-2"

>2"

20

No

10

<1%

20

20

Pages 5 and 8 Texas Coding


Guide

All Else

15

Bridge file

<5000

10

Item 29 NBI

>8"

15

Bridge file

No

15

Bridge file or Observation

Low

Degree of
Severity
Max
Score

C.9

Medium

Screening

Type of
Attributes

High

Remote

Similar Items in
Guideline

Attributes

Bridge/Deck/Cracking
Source of data

Clear Cover

Design

<1"

D.10

Overlay

Design

Yes

C.12

Spalling

Condition

>6%

C.10

Delamination

Condition

Yes

L.1

Mixed design
(Water)
ADTT

Loading

Poor Mix/Poor
H2O
>5000

D.20

Thickness

Design

<7"

D.19

Cold Joints
Cracking (map
dense)
Environment
Age Years of
Services

Design

Yes

D.8

C.9
L.3
D.6

Design

2%-5%

No

7"-8"

If
more
than
10%

Degree of
Severity
Max
Score

D.11

Screening

Type of
Attributes

High

Remote

Similar Items in
Guideline

Attributes

Bridge/Deck/Delamination
Source of data

Bridge file or covermeter


Item 108A NBI (See also pages 9
and 10 Texas Coding Guide)
Pages 5 and 8 Texas Coding
Guide

Condition

Yes

No

15

Page 30 Texas Coding Guide


(Deck Cracking)

Loading

Above I-20

All Else

20

Bridge file

Condition

50+

Other

15

Item 27 NBI (Year Built)

207

Low

Environment

Loading

Coast

North of I-20

All Else

20

Bridge file

Existing Section
Loss

Condition

Yes

No

20

Pages 10, 11, and 15 Texas


Coding Guide

D.18

Deck Drainage
onto
Superstructure

Design

Yes

No

10

Bridge file

C.22

Debris

Condition

Yes

No

10

C.4

Joint Leakage

Condition

Yes

No

10

D.13

Built-Up Riveted

Design

Yes

No

20

Pages 23, 25, and 30 Texas


Coding Guide
(Pack Rust)
Pages 23 and 24 Texas Coding
Guide
Bridge file

D.19

Deck Cold Joints

Design

Yes

No

15

Bridge file or Observation

D.6

Age Exposure

Design

Other

10

Item 27 NBI (Year Built)

C.21

Corrosion

Condition

50+
CR 3 or
Greater/No
Coating or
Weather Steel

Else

10

Pages 10, 11, and 15 Texas


Coding Guide

L.3

S.9

Degree of
Severity
Max
Score

Medium

Screening

Type of
Attributes

High

Similar Items in
Guideline

Remote

Attributes

Bridge/Superstructures/Sectionless
Source of data

Condition

Yes

D.3

Codes For Under


clear (Vehicle)

Design

<=15'-6"

17'-6"<H<=15'-6"

Max
Score

Existing Impacts

Degree of
Severity

C.6

Medium

Screening

Type of
Attributes

High

Source of data

No

20

Pages 33 and 34 Texas Coding


Guide

17'-6"<
Or No
Highway under
the bridge

20

Item 54B NBI


(Minimum Vertical Under
Clearance)

Low

Remote

Similar Items in
Guideline

Attributes

Bridge/Superstructures/Impact

208

Medium

Low

Detail Category

Design

E0 or E'

D or
Unknown

C or
Better

15

Bridge file or Observation

History of
Previous
Cracking that
was repaired

Condition

Yes

No

15

Page 30 Texas Coding Guide


(Steel Fatigue)

D.6

Year built

Design

Before 1975 or
Unknown

After
1985

20

Item 27 NBI (Year Built)

D.18

Skew Angle

Design

>30

<30

10

Item 34 NBI

L.1

ADTT
Active or
unmitigated
cracking due to
any cause

Loading

>5000

<5000

20

Item 29 NBI

D.17

C.18

S.7, C.19, and


C.20

Screening

1976-1984

Remote

Degree of
Severity
Max
Score

Type of
Attributes

High

Similar Items in
Guideline

Screening

Attributes

Bridge/Superstructures/Fatigue Cracking

Repair Must
be shown to
be working

Bridge/Superstructures/Fire Damage
Reason(s) for Attribute
Incidences of fire on or below a highway bridge are not
uncommon. This type of damage is most frequently caused
by vehicular accidents that result in fire, but secondary causes
such as vandalism, terrorism, or other damage initiators
should not be discounted. If fire does occur on or below a
bridge, an appropriate follow-up assessment should be conducted to determine how the fire has affected the load carry
ing capacity and the durability characteristics of the main
structural members and the deck. This assessment is typically
performed during a damage inspection immediately following the incident.
Damage to bridge components resulting from a fire is
either immediately apparent during the damage inspection, or

Source of data

Pages 30 Texas Coding Guide


Or Observation

may manifest within the first 12- to 24-month interval follow


ing the fire. Based on this observation, bridges that have experienced a fire may be screened from the reliability assessment
until an inspection, which has been conducted approximately
12 months or more after the fire, confirms that the fire has
not affected the typical durability characteristics of the bridge
components. The purpose of this screening is to ensure that
damage from the fire has not manifested after the damage
inspection.

Assessment Procedure
This attribute is scored based only on the occurrence of a
fire on or below the structure being assessed. It is assumed
that an appropriate assessment immediately following the
fire incident (i.e., damage inspection) has been performed.

Fire incident has occurred and an inspection 12 months


after the fire has not occurred

Bridge is not eligible for reliability assessment until


inspection confirms that the bridge is undamaged

There have been no incidences of fire on or below the bridge,


or inspections conducted approximately 12 months or more
after the fire have confirmed that the bridge is undamaged

Continue with procedure

209

--

Medium

Low

Load Posting

Condition

Cond Posting

Des Post

None

20

Item 41 NBI

Previous*
Overload
Damage

Condition

Yes

No

20

Bridge file

Highway
Ownership

Condition

Local

State

15

Item 22 NBI

Remote

Degree of
Severity
Max
Score

--

High

Screening

D.2

Type of
Attributes

Similar Items in
Guideline

Attributes

Bridge/Superstructures/Deflection Overload
Source of data

*Overload damages manifest in forms of settlement, rotation, and cracks.

L.3
C.8 (Corrosion
Induced Cracking)
C.1

ENV
Existing
Damage
Current
Condition

Loading

Above I-20

Condition

CS 4

Condition

Medium

Low

Remote

All else

H
CS 1

20

7 or greater

20

Item 60 NBI

between
1.5" and
2.5"

Greater
than or
equal 2.5"

20

Bridge File

10

Bridge File

20

Joints Condition Item 58 NBI


details in bridge file or items
#300 to #304 Texas Coding
Guide pages 23-24

Design

D.12

Rebar Type

Design

Uncoated

Epoxy
coated

5 or less

Joint
less

Condition

Geographical Map
(Texas Coding Guide Page 16
20)

5 or less

Cover

Joints Condition

20

CS 2

D.11

7 or greater

CS 4

Source of data

CS 3

1.5" or
Less,
Unknown

C.4

Degree of
Severity
Max
Score

High

Screening

Type of
Attributes

Similar Items in
Guideline

Attributes

Bridge/Substructures/Corrosion Damages (Spalling/Delamination/Cracking/Rust)

Low

D.21

Footing
Type

Design

Spread FTG on
soil/unknown
foundation

Drill shaft
friction pile
/etc

D.22

Subsurface
Condition

Condition

Slide zone, clay,


silt, shale, gravel

Limestone

Solid, rock

C.3

Existing
Settlement

Condition

Active (no
monitor data)

Occurred
but arrested

None

S.10

Scour
Rating

Screening

4-6

>7
Or
N

If foundation was based on


Rock/Piles we do not need
to deal with other
following attributes

Degree of
Severity
Max
Score

Medium

Screening

High

Remote

Type of
Attributes

Similar
Items in
Guideline

Attributes

Bridge/Substructures/Settlement

20

Bridge File

20

Bridge File

<3

Source of data

20

Item #405
Texas Coding
Guide on page
31
Item 113 NBI
(See also item
#407 on Texas
Coding Guide
on page 32)

210

APPENDIX C

Controlling Damage Modes for Sample Bridges

211

Table C1. Controlling damage modes for RBI analysis of bridges in Oregon (CF Case 4).

212

Table C2. Controlling damage modes for RBI analysis of bridges in Texas (CF Case 3).

211

Table C1. Controlling damage modes for RBI analysis of bridges in Oregon
(CF Case 4).

Bridge ID

Inspection
Interval
Based on
Case 4

02376B

48

07801A

24

01741B

24

07935A

48

07935B

48

17451

24

16454

24

16453
9546
00988A
01056A
9358
16873
18175

48
24
48
24
24
72
48

01895A

48

9915
8994
8896
20666

24
24
48
72

19739

48

19738

48

19284

48

Controlling Damage Mode


Rutting in deck and corrosion in substructure
Shear cracking, strand corrosion, and rebar corrosion within the span in
superstructure
Corrosion related damage modes in superstructure and substructure
Rutting in deck, shear cracking, strand corrosion, and bearing seat problems
in superstructure
Rutting in deck, shear cracking, strand corrosion, and bearing seat problems
in superstructure
Spalling in deck
Rutting in deck, shear cracking, strand corrosion, and rebar corrosion within
the span in superstructure
Strand corrosion
Strand corrosion and rebar corrosion within the span
Shear cracking in superstructure and settlement and corrosion in substructure
Most corrosion related damage modes
Strand corrosion and rebar corrosion within the span
All damage modes equal
Most damage modes in superstructure
Rebar corrosion within the span for superstructure and settlement in
substructure
Strand corrosion and rebar corrosion within the span
Rebar corrosion within the span
Cracking in deck
All damage modes equal
Rutting in deck, corrosion related damage modes in superstructure, and
settlement in substructure
Rutting and spalling in deck, corrosion related damage modes in
superstructure, and settlement in substructure
Settlement in substructure

212
Table C2. Controlling damage modes for RBI analysis of bridges in
Texas (CF Case 3).
Inspection
Interval Based on
Case 3

Controlling Damage Mode

01-139-0-0769-01-007
02-127-0-0014-03-194
02-127-0-0094-04-057
02-220-0-1068-02-058
05-152-0-0067-11-188

24
24
72
24
24

Corrosion in substructure
Fatigue cracking
All damage modes equal
Cracking in decksection loss, impact
Fatigue cracking

08-030-0-AA01-31-001

48

Deck damages and substructure

12-085-0-1911-01-003
12-102-0-0027-13-195
12-102-0-0500-03-320
15-015-0-0025-02-162
15-015-0-B064-55-001

48
48
48
48
72

18-057-0-0092-14-210

24

18-061-0-0196-01-133

24

Section loss and corrosion in substructure


All damage modes equal
All damage modes equal
All damage modes equal
All damage modes equal
All damage modes equal (screen because of pin
and hanger connection)
Punch through and cracking in deck, corrosion in
substructure

19-019-0-0610-06-162
23-141-0-0251-05-020
23-215-0-0011-07-056
24-072-0-0167-01-059

24
48
48
24

Bridge ID

Impact
Corrosion in substructure
All damage modes equal
All damage modes equal

Abbreviations and acronyms used without definitions in TRB publications:


A4A
AAAE
AASHO
AASHTO
ACINA
ACRP
ADA
APTA
ASCE
ASME
ASTM
ATA
CTAA
CTBSSP
DHS
DOE
EPA
FAA
FHWA
FMCSA
FRA
FTA
HMCRP
IEEE
ISTEA
ITE
MAP-21
NASA
NASAO
NCFRP
NCHRP
NHTSA
NTSB
PHMSA
RITA
SAE
SAFETEA-LU
TCRP
TEA-21
TRB
TSA
U.S.DOT

Airlines for America


American Association of Airport Executives
American Association of State Highway Officials
American Association of State Highway and Transportation Officials
Airports Council InternationalNorth America
Airport Cooperative Research Program
Americans with Disabilities Act
American Public Transportation Association
American Society of Civil Engineers
American Society of Mechanical Engineers
American Society for Testing and Materials
American Trucking Associations
Community Transportation Association of America
Commercial Truck and Bus Safety Synthesis Program
Department of Homeland Security
Department of Energy
Environmental Protection Agency
Federal Aviation Administration
Federal Highway Administration
Federal Motor Carrier Safety Administration
Federal Railroad Administration
Federal Transit Administration
Hazardous Materials Cooperative Research Program
Institute of Electrical and Electronics Engineers
Intermodal Surface Transportation Efficiency Act of 1991
Institute of Transportation Engineers
Moving Ahead for Progress in the 21st Century Act (2012)
National Aeronautics and Space Administration
National Association of State Aviation Officials
National Cooperative Freight Research Program
National Cooperative Highway Research Program
National Highway Traffic Safety Administration
National Transportation Safety Board
Pipeline and Hazardous Materials Safety Administration
Research and Innovative Technology Administration
Society of Automotive Engineers
Safe, Accountable, Flexible, Efficient Transportation Equity Act:
A Legacy for Users (2005)
Transit Cooperative Research Program
Transportation Equity Act for the 21st Century (1998)
Transportation Research Board
Transportation Security Administration
United States Department of Transportation

Anda mungkin juga menyukai